Sie sind auf Seite 1von 141

Integrating Distributed Resources into Electric Utility Distribution System

EPRI White Paper

Technical Report

Integrating Distributed Resources into Electric Utility Distribution Systems


EPRI White Paper
1004061

Technology Review, December 2001

EPRI Project Manager D. Herman

EPRI 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10412, Palo Alto, California 94303 USA 800.313.3774 650.855.2121 askepri@epri.com www.epri.com

DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES


THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN ACCOUNT OF WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH INSTITUTE, INC. (EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE ORGANIZATION(S) BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM: (A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I) WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL PROPERTY, OR (III) THAT THIS DOCUMENT IS SUITABLE TO ANY PARTICULAR USER'S CIRCUMSTANCE; OR (B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER (INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR SELECTION OR USE OF THIS DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT. ORGANIZATION(S) THAT PREPARED THIS DOCUMENT EPRI PEAC Corporation

ORDERING INFORMATION
Requests for copies of this report should be directed to EPRI Customer Fulfillment, 1355 Willow Way, Suite 278, Concord, CA 94520, (800) 313-3774, press 2. Electric Power Research Institute and EPRI are registered service marks of the Electric Power Research Institute, Inc. EPRI. ELECTRIFY THE WORLD is a service mark of the Electric Power Research Institute, Inc. Copyright 2001 Electric Power Research Institute, Inc. All rights reserved.

CITATIONS
This report was prepared by EPRI PEAC Corporation 942 Corridor Park Boulevard Knoxville, TN 37932 Principal Investigator P. Barker This report describes research sponsored by EPRI. The report is a corporate document that should be cited in the literature in the following manner: Integrating Distributed Resources into Electric Utility Distribution Systems: EPRI White Paper, EPRI, Palo Alto, CA: 2001. 1004061.

iii

REPORT SUMMARY

This EPRI white paper is about understanding electric power engineering issues related to integrating distributed resources (DR) into utility distribution systems. It is an overview designed for all stakeholders rather than a rigorous technical engineering guide. A major goal of the paper is to move discussion of integration issues toward solutions. Background Customer-owned electric generation is not a new phenomenon. From the early days of developing a wide-area electric power system, the value of using local fuels for on-site distributed resources and co-generating electricity with local heat recovery has been demonstrated. Integrating these systems into the transmission or distribution systems of local electric utilities has been engineered on a case-by-case basis, taking into account unique aspects of each site. Today, the emergence of promising new distributed resources technologies has raised expectations that distributed resources could become widespread. The distributed power industry is asking for uniform connection requirements, faster permitting procedures, and minimization of special studies. A U.S. government report in 2000 identified delays and difficulties in interconnection as barriers for distributed power projects. On the other hand, utility system engineers are concerned that many small generators or a single and relatively large distributed generator might adversely affect the electrical performance of distribution feeders. The idea that uniform connection requirements can cover the wide variety of technical situations has not been accepted. Cases have been sited that support both positions. Objectives To educate regulators/policy makers, DR developers, and other stakeholders on the complexities of integrating DR into the distribution system and to translate the technical language of DR experts into language that all stakeholders can understand. Approach The project team looked at technical questions regarding the safe and effective installation of distributed resources. Team members started with a fundamental description of electrical distribution systems, then addressed issues that arise when distributed resources are added to the system. They addressed tradeoffs between maintaining simple and uniform connection requirements and achieving a safe and effective interface in the wide variety of connection possibilities. Only part of the teams effort was to explain the grids inherent problems (for example, the grid was not originally designed for two-way power flow). They also worked to send a positive message that the utility industry is working to make DR integration as simple as possible for DR developers while preserving grid reliability and performance standards. v

Results Better understanding of technical issues assists regulators and others responsible for setting public policy, rules, and standards for interconnection at state and local levels. Though technical issues explained in the report also assist technically trained decision makers, they have been written in terms that all stakeholders can understand. These explanations serve as a tool for utility planners, energy providers, and system designers to balance needs of distributed resources with safety and reliability needs of the public power supply. Most of the technical issues addressed are relatively complex and are expected to continue to receive some level of debate among various stakeholders. Effective compromises will come from more trials and hands-on experience with DR installations. EPRI Perspective Advances in DR technologies and restructuring of the electric utility industry are simultaneously fostering increased interest in DR. When effectively integrated into an electric power system, DR can be used to provide energy, capacity, and various ancillary services such as voltage regulation, power quality improvement, and emergency power supply. Misapplication of DR can adversely impact both the private user and the public power supply. Understanding the issues will enable stakeholders to effectively address interconnection and system integration concerns. This white paper will help all stakeholders involved in DR applications discuss and address potential issues in the system design stage. Keywords Distributed resources Integration Distribution impacts Interconnection rules Distributed generation

vi

CONTENTS

1 EXECUTIVE SUMMARY ..................................................................................................... 1-1 Overview and Key Issues ................................................................................................... 1-1 Power System Compatibility With DR................................................................................. 1-3 The Probability of DR Affecting the Public Power System .................................................. 1-3 DR Integration Solutions..................................................................................................... 1-4 2 INTRODUCTION ................................................................................................................. 2-1 Focus of This Report .......................................................................................................... 2-1 Distributed Resources Technologies .................................................................................. 2-2 Value of Distributed Resources to the Transmission and Distribution System .................... 2-5 Impact of Distributed Resource Ownership on Interconnection........................................... 2-7 Interconnection Costs......................................................................................................... 2-8 The Compatibility of Distributed Resources and Distribution Systems ................................ 2-9 Anatomy of a Distributed Resource ...................................................................................2-11 Role of the Utility System Design ......................................................................................2-13 Summary...........................................................................................................................2-14 3 THE DISTRIBUTION SYSTEM ............................................................................................ 3-1 Distribution System Topologies .......................................................................................... 3-1 Radial Systems .................................................................................................................. 3-2 Voltage Levels and Capacity of Radial Distribution Systems .............................................. 3-4 Faults and Protective Devices on Radial Systems.............................................................. 3-4 Networks ............................................................................................................................ 3-8 Voltage Regulation of Power Distribution Systems ............................................................3-11 4 THE IMPACTS OF DISTRIBUTED RESOURCES ON DISTRIBUTION SYSTEMS............. 4-1 Introduction ........................................................................................................................ 4-1 Data Needed to Evaluate DR Impacts ................................................................................ 4-1

vii

Sensitivity of Power System to Location of Distributed Resources...................................... 4-2 Distributed Resources at the Distribution Substation .......................................................... 4-4 Distributed Resources on the Primary Feeder and Secondary Lines .................................. 4-7 Losses and Reactive Power Flow......................................................................................4-11 Impact of Power Factor on Losses ....................................................................................4-12 Current Ratings and Capacity Issues ................................................................................4-12 Safety................................................................................................................................4-14 Impact on Fault Conditions................................................................................................4-15 Transient Recovery Voltage..........................................................................................4-15 Fault Levels..............................................................................................................4-15 Impact of Fault Contributions of Distributed Resources on Lateral Fusing Practice ...........4-19 Impact of Reverse Flow Fault Contributions on Sectionalizers, Reclosers, and Circuit Breakers............................................................................................................................4-21 Size Limits Where Distributed Resource Fault Contributions Become an Issue.................4-23 Grounding Compatibility ....................................................................................................4-24 Ground Fault Overvoltage on a Four-Wire System .......................................................4-30 Small Generators and Ground Fault Overvoltages........................................................4-32 Transformers and Grounding ........................................................................................4-33 Use of Existing, Three-Phase Transformer ...................................................................4-34 Transformers for Single-Phase Distributed Resources..................................................4-34 Circulating Currents in Transformer Windings...............................................................4-35 Impact of Grounded, Distributed Resource on Feeder Fault Sensing ................................4-36 Single-Phasing of Transformer Banks with Grounded-Wye and Delta Connections......4-39 Transformer and Grounding Issues that Impact the Distributed Resource Itself ............4-40 Overall Conclusions on Transformers and Grounding...................................................4-41 Service Reliability ..............................................................................................................4-43 Islanding of Distributed Resources on Radial Systems......................................................4-43 Intentional Islanding......................................................................................................4-49 Avoiding Islanding on Network Systems and Interactions with Network Protectors .......4-50 Voltage and Frequency Control .........................................................................................4-55 Stability .............................................................................................................................4-60 Sags, Swells, and Momentary Interruptions.......................................................................4-61 Harmonics .........................................................................................................................4-62 Overvoltages .....................................................................................................................4-65 Flicker ...............................................................................................................................4-65

viii

5 SUMMARY AND INTERCONNECTION SOLUTIONS......................................................... 5-1 Screening of Potential DR Interconnections ....................................................................... 5-1 Standardized Interconnection Requirements ...................................................................... 5-4 Interconnection Equipment at a Typical Site....................................................................... 5-5 Relays for Protection of Synchronous and Induction Generators.......................................5-11 Relays for Protection of Inverters ......................................................................................5-13 Relay Functions.................................................................................................................5-14 Suitable Voltage and Frequency Operating Window..........................................................5-15 Synchronization.................................................................................................................5-16 Non-Exporting DR as an Alternative to Employing Protection Functions ...........................5-17 Fast-Acting Directional Protection .....................................................................................5-18 Communication .................................................................................................................5-19 Overall Conclusions on the Complexity of Interconnection ................................................5-20

ix

LIST OF FIGURES
Figure 2-1 Example of Installation Using Reciprocating Engines and Combining the Heat and Power Functions (Courtesy Caterpillar Corp.)........................................................... 2-3 Figure 2-2 Example of Building-Integrated Photovoltaic System (left) and Large-Scale Photovoltaic Concentrator System (Right) (Photographs Courtesy of NREL) ................. 2-4 Figure 2-3 Example of a Microturbine With a Rating of Approximately 28 kW (Photograph Courtesy of Capstone) ................................................................................ 2-4 Figure 2-4 Example of 200-kW Fuel Cell at Sunstrand Data Center in Windsor Locks, Conn. (Photograph Courtesy of International Fuel Cell, Inc) ............................................ 2-5 Figure 2-5 A Distribution System of the Future Showing a Wide Range of Interconnected Generation Devices........................................................................................................2-10 Figure 2-6 Interconnection of Distributed Resources and the Utility System Is Defined by the Type of Prime Energy Source (Top Row), the Type of Power Conversion (Middle Row) and the Utility System Interface and Protection Equipment in Use (Bottom Row) .................................................................................................................2-12 Figure 3-1 Examples of Distributed Resources Interconnected With the Power System. The Shaded Region Shows Where Distributed Generation Might be Interfaced. ............. 3-1 Figure 3-2 Classic Radial Distribution System......................................................................... 3-3 Figure 3-3 A Radial Distribution System Using the Auto-Loop Concept .................................. 3-3 Figure 3-4 Low-Voltage Spot Network Configuration (Note: Arrows Show the Normal Direction of Power Flow.) ................................................................................................ 3-8 Figure 3-5 Portion of a Low-Voltage Grid Network .................................................................. 3-9 Figure 3-6 Example of Network Protector and Primary Feeder Circuit Breaker Isolating a Faulted Primary Cable (Note: Arrows Show Direction of Power Flow to Fault.) ..............3-11 Figure 3-7 Voltage Regulation Devices Used on Distribution Systems Include Load-Tap-Changing Transformers, Feeder Regulators, and Switched Capacitors .........3-12 Figure 4-1 Placement of Distributed Resource at the Substation Will Generally Impact the Feeder Less Because of the Lower Impedance of the Substation ............................. 4-6 Figure 4-2 Voltage Profiles With and Without a Large Amount of Distributed Resources Connected at the Substation ........................................................................................... 4-6 Figure 4-3 Fault on an Adjacent Circuit Causes a Nuisance Trip of a Distribution Feeder because of the Backfeed of Fault Current from Distributed Resources (G1, G2, and G3) .................................................................................................................................. 4-7 Figure 4-4 Evaluation Points Where the Stiffness Ratio Should Be Evaluated ...................... 4-9 Figure 4-5 Optimal Location for a Distributed Resource to Reduce Feeder Losses................4-11

xi

Figure 4-6 An Example of Fault Contributions from Distributed Resources Being Added to the Contribution from the Substation Source and Increasing the Fault Level on the Lateral ......................................................................................................................4-16 Figure 4-7 Contribution of Fault Current From Typical Rotating Generators During the Sub-Transient, Transient, and Steady State Time Frames .............................................4-17 Figure 4-8 An Example of How Fuse-Saving Coordination Is Impacted by Use of Distributed Resources ....................................................................................................4-20 Figure 4-9 Distributed Resource Located on the Branch Circuit Contributes Fault Current Causing Confusion to Sectionalizing Switch C ...............................................................4-22 Figure 4-10 Example of Nuisance Tripping on the Feeder With Distributed Resource............4-23 Figure 4-11 Grounding of a Four-Wire, Multi-Grounded-Neutral Distribution System .............4-25 Figure 4-12 Example of the Three-Wire, Ungrounded Primary System ..................................4-26 Figure 4-13 Example of Effective Grounding..........................................................................4-28 Figure 4-14 Example of a Three-Wire, 13.8-kV Ungrounded System During a Fault to Earth ..............................................................................................................................4-30 Figure 4-15 Example of a Four-Wire System Experiencing a Fault The Overvoltage on the Distribution Circuit Is Fed by an Ungrounded Generator (With Respect to Utility Primary) During a Ground Fault. Note That the Generator Ground Connection on the Wye Side of the Transformer Does Not Make the Delta on the High-Voltage Primary Side Appear to Be Grounded. ...........................................................................4-31 Figure 4-16 Use of a Ground-Fault-Overvoltage Detection Scheme to Trip an Ungrounded Generator and Limit the Duration of Overvoltages .....................................4-32 Figure 4-17 Four Commonly Used Transformer Arrangements and the Way They Appear From a Grounding Perspective to the Primary Utility Distribution System...........4-33 Figure 4-18 A Typical, Small (<10 kW) Inverter Installation With Proper Grounding Serving a Residential Load.............................................................................................4-35 Figure 4-19 Circulating Current in Delta Winding Because of Zero-Sequence Voltage...........4-36 Figure 4-20 Neutral Grounding Impedance Used to Limit Circulating Current and Ground Fault Desensitization and to Reduce Ground Fault Contributions From the DR Itself .....4-37 Figure 4-21 A Common Transformer Arrangement Found at Many Commercial or Industrial Distributed Resource Sites..............................................................................4-38 Figure 4-22 Grounding Impedance at the Neutral Terminal for the Rotating Generator..........4-39 Figure 4-23 A Delta Winding on the Low Side Can Result (1) in the Distributed Resource Unit Back-Feed of Two De-Energized Phases, (2) in the Transformer Overloading, and (3) in Improper/Unsafe Voltage Conditions and Phase Relationships on the Feeder............................................................................................................................4-40 Figure 4-24 Example of a Large Island That Might Be Created if a Large Distributed Resource (e.g., 3 MW) Is Isolated With a Major Portion of a Distribution Feeder............4-44 Figure 4-25 Example of a Small, Unintentional Island on the Secondary System if the Fuse Should Operate or Be Removed and the Anti-Islanding Protection in the Local Generators Does Not Work Properly ..............................................................................4-45 Figure 4-26 A Distributed Resource With Improper or Ineffective Anti-Island Protection Could Pose a Threat to the Public and Utility Workers by Back-Feeding Downed Conductors.....................................................................................................................4-46

xii

Figure 4-27 Illustration of the Concept Used in IEEE Standard P1547 ...................................4-47 Figure 4-28 Anti-Island Protection Can Be Arranged as Shown Here to Isolate the Distributed Resource From the Utility System But Continue to Operate a Local Customer Facility............................................................................................................4-49 Figure 4-29 An Example of Intentional Islanding Where a DR Has Been Strategically Placed Downstream of a Recloser..................................................................................4-50 Figure 4-30 Low-Voltage, Spot Network Under Normal Operating Conditions........................4-51 Figure 4-31 Low-Voltage Spot Network With One of the Primary Feeder Cables Faulted ......4-52 Figure 4-32 Example of Distributed Resource With a Capacity Large Enough to Exceed the Network Load ...........................................................................................................4-53 Figure 4-33 Voltage Profile on a Distribution Circuit With and Without Distributed Resources ......................................................................................................................4-57 Figure 4-34 Voltage Profile on a Distribution Feeder with Distributed Resource Added near a Voltage Regulator Station....................................................................................4-58 Figure 4-35 Injected Power From a Distributed Resource at One Customer May Influence Others on the Same Secondary ......................................................................4-59 Figure 4-36 Comparison of Ideal Sinusoidal Wave and a Severely Distorted Wave ...............4-63 Figure 4-37 GE Flicker Curves...............................................................................................4-66 Figure 5-1 The Basic Screening Approach behind all DR Screening Applications................... 5-2 Figure 5-2 Example of California Rule 21 Screening Process ................................................. 5-3 Figure 5-3 Curves Identifying the System Impacts of DR Connecting to Various Points on the Distribution System............................................................................................... 5-4 Figure 5-4 Key Elements of an Interconnection....................................................................... 5-6 Figure 5-5 Example of a Fuel Cell Interface at a Residential Location..................................... 5-7 Figure 5-6 Interconnection of a 250-kVA Synchronous Generator (Alternate Transformer Shown Is for Three-Wire Ungrounded Feeders) .............................................................. 5-8 Figure 5-7 Example of an Interconnection for a Large Three-Phase Inverter .......................... 5-9 Figure 5-8 Utility-Grade Relays (Photo Courtesy of GE) ........................................................5-12 Figure 5-9 Example of an Inverter Product Intended for Photovoltaic Applications That Has Internal Microprocessor-Based Protective Functions (Photo Courtesy of Xantrex)..........................................................................................................................5-13 Figure 5-10 When the Inverter Feeds a Bus at a Different Location Than the Intended Interconnection Isolation Point, a Separate Control Package May Be Needed for That Point.......................................................................................................................5-14 Figure 5-11 Two Methods of Relatively Fast-Acting Reverse Flow Protection........................5-18 Figure 5-12 Example of Transfer Trip Used to Ensure that the DR Is No Longer Operating .......................................................................................................................5-20

xiii

LIST OF TABLES
Table 2-1 Impact of Distributed Resource Application and Ownership on the Determination of Interconnection Requirements by the Distribution Company................. 2-8 Table 4-1 Grounding Recommendations for Distributed Resources (DR)...............................4-27 Table 4-2 Recommended Transformer Configurations for Three-Phase Distributed Resource Installations ....................................................................................................4-42 Table 4-3 Harmonic Distortion Standards From IEEE 519-1992 ............................................4-64 Table 5-1 Comparison of Typical Interconnection Equipment Used in DR Applications..........5-10 Table 5-2 Protection Functions Commonly Employed On Larger Generators ........................5-15 Table 5-3 Recommended Voltage Operating Limits for Small Distributed Resources (< 30 kW) as Described in IEEE 1547 Draft 8 .....................................................................5-16 Table 5-4 Synchronization Requirements of DR at Moment of Connection ............................5-17 Table 5-5 Comparison of the Benefits of Various Reverse Current Tripping Techniques from the Perspective of the Impacts of Generator Fault Contributions on Utility System Protection Devices.............................................................................................5-19

xv

1
EXECUTIVE SUMMARY
Overview and Key Issues
This report serves as a tool to assist utility planners, energy providers, regulators, DR manufacturers and system integrators in formulating the appropriate policies and technical requirements for DR integration into power systems. It provides an overview of the technical issues associated with interconnection of DR to the power system including the protective relaying functions needed, settings of those functions, the system grounding requirements, the interface transformer requirements and many other design issues. In addition, it covers a broader context of power system impact issues such as determining how the introduction of DR will impact the voltage control, reliability, power quality and various dynamic and transient operating characteristics of the power system. The need to formulate good DR integration policies is stronger than ever. Recent advances in internal combustion engines, combustion turbines, fuel cells, microturbines, photovoltaic and wind energy sources are creating significant new opportunities for the interconnection of multiple, diverse DR to the power system. There is a great opportunity to utilize new and emerging distributed generation and power conditioning technologies to improve power system reliability, efficiency, power-quality, alleviate system constraints and reduce energy costs. Given these potential benefits and a growing momentum toward DR, it is expected that a considerable amount of DR will be interconnected to the power system in the coming years. Many stakeholders may benefit from the introduction of DR to the power system. Electric customers may achieve a lower cost of power and improved reliability. Electric utility wires companies may use DR to facilitate deferments in T&D infrastructure, reduce power system losses, and enhance system reliability. Generation companies may elect to add renewable energy to their portfolio where it can offer emissions credits, fuel security, and enhanced marketing value. Energy service companies may install DR at customer sites and sell services such as reliability and heat (cogeneration) along with traditional electricity to create a new revenue stream. Finally, the society as a whole can benefit by having a less centralized power system that is more resistant to natural and man-made disasters. For DR to realize this potential it must be interconnected and managed in a fashion that is compatible with the power system. Failure to properly interconnect and address the key system interactions can reduce the value of DR and potentially cause reliability and safety issues for the utility system and electric customer facilities. Examples of problems that have occurred include failures and subsequent fire/destruction of low voltage network protectors caused by the operation of distributed generators in parallel with a low voltage network. There are examples of 1-1

Executive Summary

utilities around the country reporting difficulty in maintaining proper voltage on radial distribution circuits with larger fluctuating DR power sources connected such as large wind turbines. Interaction of these sources with voltage regulators and the system impedance has on various occasions resulted in sustained low and high voltage outside normal operating limits, and varying voltage that can cause noticeable flickering of lights. It is also well documented that an improperly grounded generator source, feeding into a distribution circuit, can create sufficient feeder voltage to damage lightning arresters and customer loads during fault conditions so proper grounding is a key objective in any interconnection project. Fortunately, these problems and others are all completely solvable by using appropriate system design, control and protection. This report focuses on a number of key issues related to system design, control and protection including: 1. DR Protection: Use of appropriate protective functions (relays) and settings to insure the ability of the DR to respond appropriately to abnormal conditions on the power system or a problem within the DR itself. The functions are used to trip the generator off line so as to prevent unintentional islanding, sustained fault conditions, steady state operating conditions outside acceptable voltage and frequency operating limits, insure proper synchronization and proper reconnection following a system disturbance. 2. DR Grounding and Transformer Interface: DR should be correctly grounded and use the appropriate transformer interface to insure that it will be compatible with the grounding scheme employed on the power system and at the customer site. This is to insure that overvoltages and other dangerous conditions are not created during faults that could impact both the power system and customer site. 3. Managing DR Impacts on the Power System: The interactions of DR with the power system include voltage regulation equipment, overcurrent devices, transformers, capacitors and other system equipment/controls. The effect on the system can be predicted by well-understood analytical techniques and managed through the use of appropriate design of the DR, and design changes to the distribution system where needed. 4. T&D Support: DR can offer significant transmission and distribution system benefits such as relieved congestion, reduced system losses, deferment of new infrastructure, enhanced voltage regulation, etc. These benefits may or may not be achieved depending on conditions. A key integration objective is to evaluate DR from a planning perspective and determine the value it offers (either negative or positive) in supporting the T&D system infrastructure. This evaluation can be used to define strategies and equipment needs on the power system to insure safe, reliable operation and obtain positive benefits from DR on a system basis. Key aspects of this would be locating DR at the most ideal sites and controlling it in a fashion that helps reduce peak loading on the power system. The first three items listed above are the current focus of most interconnection/system integration efforts for DR. The fourth item is also important now, but in the future, is expected to receive even more attention as the electric power industry begins to look more at strategies that see the potential for a significant penetration of DR on the power system. This report attempts to bring into focus all of these issues by describing the various interactions that can occur between DR equipment and the power system, and how problematic interactions can be avoided. 1-2

Executive Summary

Power System Compatibility With DR


The power system that is in place today was not designed with distributed generation in mind. As a result, placement of generation at this level can cause a variety of undesirable conditions that include power quality problems, degradation of system reliability, reduced efficiency, potentially damaging overvoltages, and various safety issues. On the other hand, most distribution systems are well enough designed and sufficiently large in capacity, that despite the fact that they werent intended for DR, they still can handle some amount of DR as long as the appropriate grounding, transformers and protection functions are utilized by the DR. Furthermore, when such DR is added in modest quantities and operated at the right times and locations, it actually can improve the performance of the distribution system, rather than degrade it. As DR penetration grows beyond just modest levels and up to a significant fraction of the total distribution system load, then some of the incompatibility issues of distribution system designs become very apparent and changes to the distribution system equipment and protection are needed to facilitate successful integration of DR. For example, voltage regulator controllers may be severely confused by reverse power flows caused by DR and power system protection devices (circuit breakers, reclosers, sectionalizers and fuses) may incorrectly operate when enough DR is added to increase the fault current level significantly. Fortunately, the various system impact problems are all completely solvable with available technologies that can be used to upgrade the feeders. These upgrades may involve the addition of protective relays at the substation, new feeder protection gear, changes to the voltage regulation equipment, upgraded switchgear, revised system grounding, transfer trips, and a host of other possibilities. The point is that system changes can be made to make DR work if a problem is identified.

The Probability of DR Affecting the Public Power System


Many of the more serious feeder-wide problems that can result from the installation of DR are related to large aggregations of DR or larger individual DR. A single, small fuel cell cannot cause a significant impact on a distribution system primary, but it may cause problems at the secondary level, adversely impacting power to a few houses. The probability for adverse effects increases if the DR is located on a weaker-than-average secondary circuit, and if it does not have proper protection and interface equipment. There is no general rule of thumb for the size where system impacts become an issue because so many factors play a role. These factors include the impedance (strength) of the distribution system at the point of connection, the type of distribution system design, and the characteristics of the DR (size, fault contribution rating, harmonic output, etc.). A very strong interconnection point on a 34.5 kV primary distribution feeder can tolerate many megawatts of connected DR without there being a significant impact on the primary system. On the other hand, a weak interconnection point at the end of a 13.2 kV rural distribution feeder may not be able to handle more than a few hundred kilowatts without significant impacts on the primary feeder. On the secondary side of the system, the effects are more pronounced due to the limited capacity of 1-3

Executive Summary

secondary systems. For example, a DR of just 15 kW could substantially impact a weak secondary circuit and its adjacent customers. Overall, any DR, even a small one, can cause significant impacts locally for the DR site and/or adjacent customers. As the size of the DR grows, the impacted region of the power system grows until the entire distribution circuit is affected for very large units or large aggregations of DR. The likelihood of problems occurring can be accurately predicted based upon the design of the power system and types/sizes of DR connected. Tools that enable DR system designers, owners, and operators to predict the compatibility of DR installations are both emerging and are already available, including various power system analysis software packages that can analyze the effect of DR on fault levels, voltage levels, harmonics, stability and other factors. Screening tools and methodologies to efficiently and quickly determine the DR applications that are most likely to pose a risk to the system performance are crucial to integrating large quantities of DR to the power system. These screening tools are just emerging and beginning to be used in industry and should facilitate a more rapid and streamlined process for identifying system impacts and interconnection requirements for DR.

DR Integration Solutions
Solutions to interconnection complexity are forthcoming, but efforts to improve available interconnection standards, practices, and products must be accelerated. Public policy decisions can help in promoting these improvements. To facilitate the interconnection of DR, the stakeholdersin a collaborative effortmust undertake the following actions: Develop sound technical interconnection standards that enable as close as possible plugand-play installation of DR without sacrificing safety or causing significant interference with the operation of the public power system. Establish a certification program for DR equipmentsimilar to UL ratingsto make DR behavior during various system conditions as predictable as possible. Adopt planning and design practices for future distribution systems that facilitate easier and safer interface of DR and evolve a public power system that is more compatible with DR. Encourage and provide incentives for the use of DR to support the existing transmission and distribution system. Innovate new types of fast-acting relays and power-conditioning devices to mitigate existing interconnection problems such as fault contributions from generators, overvoltages due to poor grounding, etc. Develop screening methodologies that allow efficient determination of the requirements for interconnection and system impacts of DR.

In addition to these activities, there is no substitute for experience. Stakeholders are gathering experience on the interconnection of DR to the system, but they need more experience in the behavior of multiple DR units interacting on the same distribution circuit and the response of power-electronics-based converters. Continuing field experience (as more units are brought online) and field experiments will provide case studies and data that can be used to predict and 1-4

Executive Summary

thereby prevent interconnection problems. Such information can inform simulations and laboratory studies of operating equipment to extend knowledge even further. The cost of implementing solutions to interconnection problems can be relatively low if DR-compliant system designs and interconnection techniques mature before distributed generators are installed en masse. Accelerating the pace of DR interconnection is certainly possible, but may prove more costly, more difficult, and more time-consuming if premature applications overleap designs. From a technology perspective, a fully integrated, highly DR-dependent public power system is completely possible by employing the proper integrated protection and control systems, and by redesigning distribution systems. But stakeholders must first address essential questions. Is an accelerated pace of interconnection an efficient investment of private and public resources? Will DR technologies develop to where market pressures and public policy favor an accelerated approach?

1-5

2
INTRODUCTION
Focus of This Report
This report serves as a tool to assist utility planners, energy providers, regulators, and other distributed resource (DR) stakeholders in formulating the appropriate policies and technical requirements that balance the needs of distributed generation with the safety and reliability needs of electric power systems. The use of DRs represents a fundamental change from the traditional model of centralized electric power generation and delivery. Defined as small, modular electric generation or storage devices deployed near the point of consumption, DR can either be integrated with the grid or operate independently. No universally accepted size breakpoint defines DR versus traditional, centralized generation. However, the rated capacity of most DR is generally less than 10 MW and some units, such as small photovoltaic systems, can be as small as just a few hundred watts. DR is interconnected at the distribution system or customer use levels and, depending on local regulatory issues, may be owned/operated by electric customers, power generation companies, utility wires companies, and energy service companies. Much of the current focus is on electric customer ownership of DR. However, in the future, use of DR by all parties mentioned above is likely to increase and the broad perspective is to consider how this generation, regardless of ownership, can be successfully integrated into the power system. Potentially, distributed resources offer many benefits over the traditional, centralized power system model. These benefits may include reduced cost, improved efficiency, and improved reliability. Conversely, serious difficulties can exist in integrating DR into the power system because it was never designed with DR in mind. These difficulties can only be avoided through careful planning, engineering and good regulatory policy. The material is presented in four Chapters that include: Chapter 2, Introduction (A general introduction to DR integration issues) Chapter 3, The Distribution System (A discussion of power distribution system structure and components as they relate to DR integration issues) Chapter 4, The Impacts and Distributed Resources on Distribution Systems (A discussion of the system impacts and DR equipment issues that must be considered when integrating DR to the distribution system)

2-1

Introduction

Chapter 5, Summary and Interconnection Solutions (A discussion that addresses interconnection and provides examples of how DR can safely and practically be integrated with the grid, as well as ongoing efforts to adopt uniform standards)

The information provided in this document assumes some familiarity with, but not necessarily expertise in, the distribution system and related issues.

Distributed Resources Technologies


A wide-range of power generation technologies are currently in use or under development for use in the role of DR. These technologies include: Reciprocating engines Small combustion turbines and microturbines Small steam turbines Fuel cells Small-scale, hydroelectric power Photovoltaics Solar-thermal electric Wind turbines Stirling engine systems Energy storage technologies

The above technologies may be configured in sizes ranging from residential scale up to many megawatts and can be installed just about anywhere on the distribution system. DR technologies have reached various stages of development. Some technologies, such as steam turbines and reciprocating engines, are quite mature having an experience base approaching one century, and others, such as fuel cells, are in the early commercialization phase and are still experiencing the technology development necessary for widespread economical application. Nationwide, there are already many thousands of significantly sized (greater than 1 MW) reciprocating engine and turbine installations currently operating that could be classified as DR. Typical applications include hospitals, university and college campuses, commercial buildings, wastewater treatment facilities, incinerator facilities, government buildings and manufacturing facilities. Figure 2-1 is an example of a typical, reciprocating enginebased, cogeneration system used for a college campus. Strong recent growth in the number of these installations reflects the confidence of many electric customers that reciprocating engines or small turbines operating in parallel with the utility can be an economical solution in a variety of applications. When properly equipped, DR plants can be switched into a non-parallel (off-grid) configuration suitable for emergency power during utility system outages. Manufacturers of reciprocating engine and turbine generators have developed 2-2

Introduction

modular packages that can be configured to provide these functions and operate at costs that are competitive with utility rates in many areas.

Figure 2-1 Example of Installation Using Reciprocating Engines and Combining the Heat and Power Functions (Courtesy Caterpillar Corp.)

Many emerging DR technologies also exist. Some examples include fuel cells, microturbines, and photovoltaics as shown in Figures 2-2, 2-3, and 2-4. These technologies are entering the early commercialization phase and expected to see continued performance enhancements and cost reductions that will likely lead to wide-scale deployment by the end of this decade (2010).

2-3

Introduction

Figure 2-2 Example of Building-Integrated Photovoltaic System (left) and Large-Scale Photovoltaic Concentrator System (Right) (Photographs Courtesy of NREL)

Figure 2-3 Example of a Microturbine With a Rating of Approximately 28 kW (Photograph Courtesy of Capstone)

2-4

Introduction

Figure 2-4 Example of 200-kW Fuel Cell at Sunstrand Data Center in Windsor Locks, Conn. (Photograph Courtesy of International Fuel Cell, Inc)

In addition to the various power generation technologies such as solar, wind and fuel cells, a likelihood also exists of increased use of energy storage systems within the next 5-10 years. Energy storage technologies are advancing rapidly and even though they produce no net energy, they are classified as distributed resources because they can dispatch stored energy into the power system for significant periods of time and can provide system support benefits similar to those of traditional power generation when used in this manner. Energy storage technologies being considered for DR include battery energy storage (BES), super-conducting, magnetic energy storage (SMES), flywheel energy storage (FES), compressed-air energy storage (CAES) and electrochemical, capacitor energy storage (often referred to as supercapacitors or ultracapacitors). These will need to be interfaced to the power system much like the generator technologies in regards to their protective relay, grounding and transformer interfaces because they are essentially generators during the time at which they discharge their stored energy back into the power system.

Value of Distributed Resources to the Transmission and Distribution System


Appropriately used distributed resources can potentially reduce the demand on the transmission and distribution system infrastructure, providing what are commonly referred to by utility 2-5

Introduction

engineers as Transmission and Distribution (T&D) system support benefits. System support benefits can include: Deferred investment in new or upgraded T&D infrastructure Reduced T&D system losses Relieved T&D system congestion Enhanced service reliability and power quality Improved voltage regulation

These potential system support benefits are achieved to varying degrees depending on the condition and state of loading of the electric power system, the locations and capacities of DR on the system, and the times at which DR is operated relative to peak system loading. Ideally, DR can provide significant support benefits with substantial economic value. Studies have shown that a kilowatt-hour of energy injected at the distribution level can have twice the value of a kilowatt-hour of energy injected at the bulk system level. This simply reflects the fact that the cost of energy delivered to customers is not only its central station generation cost, but also its delivery cost. By producing power near or at customers, it has higher value because it offsets the need to build new or upgrade existing power delivery infrastructure. While this concept is elegantly simple, it is important to recognize that this value can only be realized to the fullest extent under conditions where the DR has been carefully planned and integrated to avoid negative system impacts and is sufficiently reliable so that it can be dispatched any time that it is needed to help meet system demand. Unfortunately, poorly planned and poorly interconnected DR can actually deteriorate the performance of the power system leading to negative support benefits. For example, the reliability of the power system may be degraded if the DR is not properly coordinated with the electric power system protection (that is the operation of upstream circuit breakers, fuses and other devices). Sometimes, additional, costly T&D equipment may be required to avoid negative impacts on the power system so that the addition of DR actually increases the investment required in T&D infrastructure as opposed to deferring it. Other problems include power quality impacts ranging from poor voltage regulation to voltage flicker and harmonics that can arise from the introduction of DR. Even serious safety and equipment damage issues can exist that impact other electric customers, the general public, and utility company workers and equipment. Issues exist that utility engineers and planners need to consider every time a distributed generation device is interconnected to the system. Good safety screening and system impact-screening procedures for the interconnection of DR are critical to ensure that negative consequences can be minimized and that the positive impacts of DR can be achieved without excessive interconnection requirements for most DR. The ability to achieve T&D support benefits has as much to do with the time and power level at which the unit is operated, as it does with the technical details of the interconnection equipment. 2-6

Introduction

For example, a distributed generator could have an excellent interconnection design and all the right equipment but if it cant be relied upon to operate during the time of peak distribution system demand, then it will not help defer T&D investments. The generators operation must be coordinated with the distribution system peak demand if T&D deferment benefits are to be achieved. In the case of customer-owned DR, it may be operated at a time that provides benefit for its owner but not for the distribution system as a whole. Unless incentives are offered, commercial and industrial electric customers that install DR would not be expected to be particularly interested in operating their generation in coincidence with peak system demand. Of course, many cases exist where the customers DR usage happens to be coincidental with the utility system needs as a whole and so in those situations the unit does help reduce peak loading on the power system.

Impact of Distributed Resource Ownership on Interconnection


Electric customers, power transmission and distribution companies, power generation companies, and energy service companies may own distributed resources service companies. Much of the focus of DR integration in recent years has been on electric customerowned DR that is operated solely for the benefit of the customer. However, a rising interest also exists at many utility companies in owning and operating DR for T&D support. In this approach, utilities would install generators at strategic distribution system locations. These units would be controlled and dispatched by the utility company and would be deployed in lieu of conventional T&D system upgrades. Because the units could potentially be placed at optimal locations and dispatched with high reliability, the support benefits can be more easily achieved than with randomly placed, customer-owned DR that ordinarily is not dispatchable. Whether wires-companies, power marketers, or electric customers own DR, the system impact issues are still governed by the same physics. However, ownership variations do focus the DR applications differently and this does cause differences in the operation of the DR, and therefore its system impacts. For example, a wires-company that wants to own and operate DR will likely be looking for ideal support locations on the distribution circuits. These will be the locations that avoid negative impacts and provide the most positive system support impacts. Whereas, a customer-owned DR located at a fixed point will not have a reason or ability to move his site to the ideal feeder location. As a result, the customer may be forced to deal with resolving a negative system impact that the utility wires company would simply avoid by using a different site. Because customer-owned generation is not necessarily dispatchable at the needed times, the objective of most utilities when dealing with it is not to obtain system support benefits for T&D deferment. Rather, it is to simply avoid any serious negative system impactsin other words the utility company wants to make sure that the generator has at least a neutral impact on the distribution system. It is an added bonus if positive support of the T&D system can be obtained from a customer-owned generator. 2-7

Introduction

In yet another ownership mode, power marketers/aggregators that operate DR may wish to export more power into the distribution system than a typical customer-owned or utility wirescompanyowned DR. Such installations may require a more detailed analysis and interconnection requirements when the mode of operation encourages large exports of power into the system. Overall, these differences in application for each type of DR owner result in significant differences in the focus of interconnection studies and requirements (Table 2-1 summarizes these differences).
Table 2-1 Impact of Distributed Resource Application and Ownership on the Determination of Interconnection Requirements by the Distribution Company Type of DR Application and Ownership Mode Primary Focus of the Distribution Company in Determining Interconnection Requirements Identify any problematic distribution system impacts and safety concerns that must be corrected before the unit can operate in parallel the distribution company is generally happy if the impact on the system is neutral. It is an added bonus if a positive impact (T&D support) can be achieved. Identify any problematic system impacts and safety issues, but also to identify distribution system locations and DR interconnection requirements that can provide the most positive support of the T&D system. The focus here is definitely positive T&D support. The power market/aggregator is looking for interconnection sites that offer the greatest value and need for exported power, but the distribution company is focusing on insuring that the interconnection sites under consideration do not lead to problematic distribution system impacts and safety issues. The distribution company may be able to encourage sites that provide positive distribution system support and meet the needs of the power marketer/aggregator. The focus of the distribution company is similar to the focus they would apply for customer owned DR in most cases. By working with the energy service company, there may be limited ways that installations could be encouraged that provide the most value for distribution system support.

Electric Customer Owned DR Installed for Customer Purposes Only

Distribution Company Owned DR Installed for T&D Support or Incentives Given to Customer Owned DR for Such Support Power Marketer Independent of Distribution Company Attempting to Locate a DR on Distribution System and Sell Power to Bulk System

Energy Service Company Installing DR at Electric Customers to Serve Customer Needs

Interconnection Costs
For customers that own DR, a desire exists to minimize the cost of the utility system interface equipment associated with their generator. This includes items such as relays and controls, monitoring systems, switchgear, the interface/isolation transformer, and other equipment that are needed for the utility interface. From the electric customers perspective, these items dont necessarily provide a direct benefit for them, but they can be expensive items when properly specified according to utility system needs. This, of course, puts the utility company and DR owner at odds when it comes to the requirements for a particular DR installation. In some cases, the cost of the utility interface equipment is too expensive to justify the DR installation. 2-8

Introduction

The reasonable safety and reliability needs of the power system should always take precedence over the economic viability of any particular DR installation and so interconnection requirements can still be reasonable even when they may be prohibitively expensive for DR owners. The electric customer that owns DR usually understands this and certainly does not want an unsafe or problematic installation, and they would certainly accept any requirements that are proven technically sound and not based on achieving a higher standard of safety or quality than the utility company demands of its own equipment. However, the area where misunderstanding often occurs is that the customers attempting to install DR dont understand the utility companys motivation and need for various specific technical requirements of the interconnection. Education of the customers regarding power system operation and design issues, as well as DR integration issues, is critical to resolve the ongoing conflicts that arise. One of the key efforts underway in the power industry is the development of regulatory and technical standards that facilitate the lowest cost and fastest methods for DR interconnections that dont unduly sacrifice the safety or interfere with the operation of the power system. The industry is not there yet because the technology and technical consensus on the requirements are still evolving. Furthermore, huge investments have to be made in equipment (both DR-side and utility system side equipment). Bad judgment in regulatory policies and requirements for DR could waste huge investments and create all sorts of problems that could plague the industry for many years. All stakeholders including electric utilities, customers, DR manufacturers, and regulators must come together to carefully create a consensus set of reasonable requirements and approaches that balance the key needs of the power system and facilitate the timely introduction of DR to the system.

The Compatibility of Distributed Resources and Distribution Systems


Despite many potential advantages of DR for both the electric customer and the utility T&D system, the installation and operation of distributed resources on distribution systems as they are currently designed can represent a significant engineering challenge for power system planners and operators. Although modest quantities of DR have successfully existed on the utility system for most of its history, these installations tended to be sparsely located and larger commercial/industrial systems that could be custom-engineered, substantially protected, and monitored closely by trained personnel. However, the large variety of DR expected to be interconnected to the system in the near future may lead to situations like that portrayed in Figure 2-5. In such cases, DR may serve most of the load on the feeder and numerous DR may be operating at any given time from a variety of locations on the distribution system. Unlike the larger commercial and industrial DR of the past, many of these new DR units will have little or no technical oversight during their daily operationtheir small size would make that type of oversight economically unfeasible. Ensuring that all of this generation can function safely within the framework of common distribution system designs, becomes a challenge as the penetration level increases. 2-9

Introduction

Figure 2-5 A Distribution System of the Future Showing a Wide Range of Interconnected Generation Devices

Given the apparent robustness of the U.S. power system, many people ask why there should be any difficulty in interconnecting DR to it. After all, most of the DR is expected to be small in capacity, relative to the capacity of the distribution system, and many people hope that the application of DR can be treated like plugging in any standard appliance such as an electric dryer or washing machine. However, the power distribution systems in operation today were not originally designed to accommodate distributed resources. The voltage regulation practices, overcurrent protection approaches, switching techniques, and maintenance and power restoration practices used for distribution systems have been based on the fundamental assumption that there are no generation sources on the distribution system. Generation sources on the distribution system can create operating conditions that cannot be managed in conventional ways and often require special equipment, protection and control provisions to operate safely and effectively. With todays DR equipment and distribution system designs, it is difficult to consider DR as a simple, plug-in appliance for all but the smallest generation devices. In the future, after much evolution of the power system design, further standardization of DR equipment and experience in dealing with DR issues, the plug-in appliance approach will eventually make sense. As we move beyond the days when just an occasional commercial- or industrial-scale DR was interfaced to the system, the key DR stakeholders need to work together to ensure that the proper 2-10

Introduction

technical foundation has been constructed upon which large numbers of DR can be successfully interfaced to the power system. There potentially could be millions of generators connected to the system in the next decade. In this environment, the luxury will no longer exist of a detailed review of each installation and ongoing oversight of the operation of each DR by a plant engineer. The technical/regulatory foundation needed to support this new environment includes the following elements: Sound technical interconnection standards that come as close as possible to plug-in appliance applications without sacrificing safety or causing significant interference with the power system operation. DR equipment certification (like UL ratings) to make DR behavior under various system conditions as predictable as possible. Screening methods to quickly and easily identify the most problematic DR installations. Adoption of distribution system planning and design practices that consider DR and the application of evolutionary design changes to distribution systems to facilitate easier and safer interface of DR (in other words, the power system of the past was not designed for DRlets start updating designs now to be more compatible with DR). Encourage and provide incentives for the use of DR for T&D support applications where appropriate.

Anatomy of a Distributed Resource


A typical distributed resource and its interconnection equipment can be broken into the components shown in Figure 2-6. The first component is the prime energy source such as photovoltaic (PV), wind turbine, fuel cell, reciprocating engine, or combustion turbine. The prime energy source produces energy in a form that is not typically suitable for injection into the utility power system. For example, it may be direct current (such as PV or fuel cell) or it may be mechanical energy (such as with a rotating shaft on a turbine or reciprocating engine). Knowing the nature of the prime energy source is important because each type of source has characteristics that define how the DR as a whole interacts with the power system. For example, solar and wind sources fluctuate in output as the wind and sun conditions change. The second key component for the DR is the power converter. The power converter (sometimes called power conditioner) changes the prime energy source into suitable, 60-Hz power. The three basic categories of power converters are: Synchronous generator Induction generator Static power converter

The first two convert mechanical rotational energy into electrical power and are often called rotating power converters because they have rotational parts. The third type, the static power converter (also known as an inverter), is composed of solid-state devices such as transistors. The term static refers to the fact that the transistors have no moving parts. In static power 2-11

Introduction

converters, transistors are switched on and off to convert direct current or other unsuitable power into 60-Hz power (PV, fuel cell, and microturbines usually use static power converters). Most DR sources are sold as integrated systems with the power converters integrated with the prime energy source. For example, a gas turbine or reciprocating engine comes equipped with a matched, synchronous generator. A fuel cell or microturbine generally is equipped with an integrated static power converter. One exception is PV for which the modules that produce DC power are sold separately from the static power converter in most cases. It is important to recognize that some prime energy sources may be integrated with several different types of power converters. For example, one manufacturer of microturbines has elected to use an induction generator with their microturbine and other microturbine manufacturers have chosen to employ static power converters. Wind turbines also employ different types of converters including induction generators and static power converters depending on their design.

Figure 2-6 Interconnection of Distributed Resources and the Utility System Is Defined by the Type of Prime Energy Source (Top Row), the Type of Power Conversion (Middle Row) and the Utility System Interface and Protection Equipment in Use (Bottom Row)

The characteristics of the power converter play a big role in the impact of the DR on the distribution system and hence the interconnection requirements needed to ensure safe, reliable 2-12

Introduction

operation. Induction generators, synchronous generators, and static power converters all respond extremely differently when subjected to various power system conditions that may occur. The fault current contributions, load step response, impedance characteristics, harmonic outputs, and dynamic stability are drastically different for each converter type. The utility engineer and DR systems integrator need to know the type of power converter being used to determine the appropriate power system interface requirements and to assess system impacts. The third part of the DR system shown in Figure 2-6 is the utility system interface and protection equipment. This is the equipment between the power converter output terminals and the utility system primary. This will usually include a step-up/isolation transformer, metering, switchgear, and perhaps some additional controls and protective relaying. It may also include a communication link to a dispatch or control center. Some of this interface equipment may be owned by the utility. For example, the utility company may own the step-up/isolation transformer, because in many cases, the DR owner will use the existing utility distribution transformer feeding their site as the step-up transformer. All elementsfrom the prime energy source to the interface transformer and protective relay functionsmust be accounted for and properly coordinated with the distribution system. A design can be perfectly suitable in most respects, but one poorly chosen element, such as an incorrect transformer grounding arrangement, can lead to problems for both the utility system and its other customers.

Role of the Utility System Design


Another factor determining the requirements for DR interconnection are the characteristics of the utility distribution system. Each utility system has its own specific protection settings, grounding configuration, voltage regulation settings, ground fault detection capability, and other characteristics. The DR must be coordinated with the operation of the distribution system based on these design characteristics. Many different distribution system designs exist, including network, manual and auto-loop systems, primary and secondary selective systems, and pure radial systems. Many grounding configurations also exist, including three-wire, ungrounded neutral; four-wire, ungrounded neutral; and four-wire, multi-grounded neutral. Furthermore, significant variations exist from one part of a system to anotherfor example, some parts of a distribution system may use a four-wire, multi-grounded neutral, while other parts of the same system may use a three-wire, ungrounded neutral. Considerable variations even exist in the impedance (relative electrical strength) of the system. Near the substation, the system may be very strong (that is, having low impedance) and near the end points of the feeder, it may be weak (that is, having much higher impedance). Weak sections of the system have more difficulty accepting large-size DR without experiencing voltage control and other problems. Knowing the impedance of the system at the point of DR interconnection and knowing the type of grounding and interacting distribution system control/protection devices is critical in determining size limits and in selecting appropriate interface equipment. 2-13

Introduction

Integration approaches for DR also need to consider the aggregate impact of DR on the distribution system. As a hypothetical example, consider a power system where the interconnection of the first 500 generators each rated at 1 kW is done easily without any significant modifications to the distribution system. Upon addition of the 501st generator, a threshold could be reached where the system impacts become serious enough to warrant upgrades to handle any additional DR capacity. To manage this aggregation effect in ongoing planning efforts, utility companies will need to keep a database of the locations and sizes of various connected DR and assess the thresholds at which system changes are required.

Summary
This introduction has provided an overview of the major issues and philosophy associated with integrating DR into the power system as a preface to the more detailed discussions in Chapters 3 to 5 of this report. While the tone of this introductory section has necessarily been serious, it is not intended to discourage DR but rather to identity the need for careful planning in its integration and interconnection. DR integration is a complex issue involving the characteristics of the DR, the power system to which it is connected, and various interactions between utility system equipment and DR. DR can have an extremely positive impact on the power system and so should be encouraged by energy industry stakeholders, but it needs to be integrated in a fashion that does not degrade the safety, reliability, quality of delivered power, or efficiency of the power system.

2-14

3
THE DISTRIBUTION SYSTEM
Distribution System Topologies
The traditional electric power system delivers electricity generated at large central station power plants through a network of transmission lines, sub-transmission lines, and distribution substations and primary/secondary distribution circuits. Distributed resources will interface to the system at the distribution system level (see Figure 3-1). In this section of the report, the focus will be on the power distribution systemdescribing some of the design and operating philosophies and how they relate to the implementation of DR.

Figure 3-1 Examples of Distributed Resources Interconnected With the Power System. The Shaded Region Shows Where Distributed Generation Might be Interfaced.

Many power distribution system topologies are used in the country; however, most customers are served by one of two basic system configurations: radial and networked distribution. The 3-1

The Distribution System

majority of distribution circuits in the United States are radial, which may be subdivided into circuits of the open-loop type or pure radial designs. Whether a system is open-loop or pure radial, it still functions as a radial power flow system under normal conditions. Most rural and suburban areas are served by such radial flow distribution systems. Downtown areas of major cities are often served by low-voltage network systems. Networks can be subdivided into spot networks and grid networks. Spot networks tend to serve high-rise office buildings and grid networks serve low-rise, downtown areas of major cities. The introduction of DR impacts both network systems and radial systems. Each system is designed and operated differently and these variations must be considered whenever generation is interfaced to the system.

Radial Systems
Radial distribution systems consist of a substation acting as a hub with radial feeders emanating outward from the station. The substation will have one or more large power transformers serving one or more main buses. The total substation power rating may be anywhere from a few MVA for tiny stations up to hundreds of MVA for extremely large substations. Typical substations are in the range of 20-50 MVA of total power capacity and have two large transformers serving two separate buses. Each bus will have several feeder circuits connected to it; each protected with a circuit breaker. Under normal conditions, this feeder breaker connection is the only source of power that supplies the feeder, so power always flows outward from the substation toward customer loads on the feeder (see Figure 3-2). A key characteristic of a radial system is that because it is served by only a single source, then opening any protective device should normally de-energize the entire power system downstream of that device. There are also open-loop, radial designs, which are normally fed from a single direction; however, they can be backfed from another direction by closing of manual or auto-loop switching devices under special conditions. Figure 3-3 shows an auto-loop configuration. The purpose of the manual and auto-loop switching schemes is to provide an alternative path to serve customers if the primary path becomes faultedthis helps utilities keep reliability high.

3-2

The Distribution System

Figure 3-2 Classic Radial Distribution System

Figure 3-3 A Radial Distribution System Using the Auto-Loop Concept

3-3

The Distribution System

Voltage Levels and Capacity of Radial Distribution Systems


The nominal phase-to-phase voltage levels of most primary distribution circuits used in the United States are between 4.16 kV and 34.5 kV. Depending on the distribution system voltage rating and conductor size used, these circuits can supply anywhere from a few megawatts up to more than 40 MW each. The 15-kV-class voltage level is by far the most prevalent, comprising more than 80 percent of the distribution circuits in the United States. The term 15 kV class refers to circuits with phase-to-phase operating voltages of 12.47, 13.2, 13.8, and 14.4 kV. The 15 kV class circuits typically have trunk feeder capacities on the order of 8-12 MW depending on conductor size. These circuits typically have three-phase, four-wire primary feeders (three-phase with a multi-grounded neutral conductor) or three-wire, ungrounded feeders. Primary feeders are generally between 5 and 25 kilometers (3-16 miles) long and form the main line from which various three-phase branches and single-phase lateral lines split. Small distribution transformers located on poles or concrete pads near customers serve anywhere from one to 20 customers each. They step down the voltage from the primary feeder level to secondary service levels in the range of 120 to 600 volts. For radial systems, these secondary circuits are rarely more than 200 meters long because of voltage drop limitations. Because they lack redundancy, radial systems are less reliable than networked systems. Despite the fact that they are among the most unreliable distribution configuration used, most open-loop, radial systems in the United States still have an availability that is better than 0.9997! The term availability is used to refer to the amount of time the system is operating and 0.9997 means that less than 3 hours of interruption time occurs each year. On radial distribution systems, the real power flow (watts) at any point on the primary feeder is always in one direction: from the substation outward. The continuous, outward flow of power implies that the current at the ends of feeders is less than it is near the substation. Just as tree branches get smaller as you move away from the tree trunk, the primary feeder divides its power up among the various branches and lateral lines. The radial power on distribution systems has a beneficial impact on the protection and voltage regulation design of the distribution system. Radial flow allows faults to be effectively cleared with simple protection schemes that do not need directional- or impedance-based relays. In other words, the fuse and breakers simply trip or burn open, based on the magnitude of the current as opposed to discriminating the location and direction of the fault current. Without radial flow patterns, the distribution system would need directional or impedance zone-based relaying. These are more complex relays that can detect the direction of current or the approximate location of the fault.

Faults and Protective Devices on Radial Systems


A failure in the insulation of a line is referred to as a fault. Faults are short circuits that have extremely high current compared to normal, operating currents on lines. When a fault occurs, the faulted section of line or equipment needs to be de-energized as quickly as possible to avoid 3-4

The Distribution System

further damage to the faulted device, damage to other power system equipment, the melting of wires and cables, and exposure of the public to various dangers associated with fault current. Wind, tree contacts, car accidents, animals, lightning, accidental cable dig-ins, and various other factors cause faults. Fault protection equipment on the distribution system helps ensure both safety and reliability by minimizing the duration of the faults and the area of the power system affected. When a fault occurs, protection equipment will usually de-energize the area affected within a few seconds or much less (can be less than 1/10th of a second). The locations of fault protection equipment and control settings determines how many customers will lose power in the event of a fault, and to a certain extent, how quickly power will be restored. Fault protection is generally based on dealing with two types of faults: temporary faults and permanent faults. Temporary faults, such as lightning flashovers, can usually be cleared or eliminated by de-energizing the circuit and restoring service (reclosing) a moment later. Permanent faults, such as hard permanent contact between conductors, cannot be cleared by a simple process of de-energizing the power and must be fixed before restoring electric service on the affected line section. Studies show that most system faults are temporary, so most utilities strive to clear the temporary fault using automatic circuit reclosing. Devices called circuit breakers and reclosers may employ the reclosing technique (discussed in detail later). Faults attributable to lightning, animals, and sometimes tree limb contact are often temporary and are best dealt with using reclosing. In a radial distribution system, fault protection is provided using devices such as fuses, electromechanical and solid-state, time-overcurrent relays that control circuit breakers and hydraulic or electronically controlled reclosers and sectionalizers. These elements are cascaded across the system to protect various sections. Their operation must be closely coordinated to ensure that faults are properly detected and cleared from the system. Circuit breakers are usually used at the substation to protect each of the main feeders emanating from the station. The sensitivity and tripping characteristics of the circuit breaker are determined by the relays used to control it and the settings employed in those relays. Fuses are used to protect small sections of the feeder such as single-phase lateral branches or occasionally three-phase branches and so are usually located out on poles some distance from the station. Fuses are the simplest type of fault protection equipment and function by melting, and thus interrupting current flow when they experience excessive current. Utilities coordinate fuses with other protective devices to ensure that during a fault the appropriate device responsible for clearing the fault will actually perform that task. The radial nature of power flow makes this coordination easier than it would otherwise be. Many utilities employ fuse saving as a practice. This approach allows temporary faults on lateral branches to be cleared by circuit breakers or reclosers and permanent faults to be cleared by fuses. Careful coordination between the time-current clearing curves of the fuse versus the tripping characteristics of upstream circuit breakers and/or reclosers is required to achieve this effect. Later in Chapter 4 of this report, we will discuss how the introduction of DR can confuse these types of protection schemes.

3-5

The Distribution System

Automatic line sectionalizers are another type of protective device that are essentially motor-operated switches that work in concert with reclosers and other protective elements to isolate faulted sections of the distribution line and thereby confine outages to small sections of the distribution line. They are not designed to open during faults but will open after an upstream device such as a fuse or circuit breaker has cleared fault current. They are actuated by fault-counting relays that track the number of reclosing cycles made by the recloser or circuit breaker, and just before the final cycle that will lock the recloser out permanently, the relay trips the sectionalizer. The sectionalizer locks open and prevents the faulted section from being re-energized the final time. Then the recloser or circuit breaker closes to restore service to all points up to the sectionalizer. Sectionalizers can also be used in auto-loop schemes to backfeed a section of a distribution system when the normal path becomes blocked because of fault or other problems. Again, the presence of DR in sufficient quantities can confuse these types of protection and control schemes, and this will be discussed in detail in Chapter 4. Reclosers are another type of protective device. Reclosers function a lot like the feeder circuit breakers found in substations, but they are self-contained units suitable for pole-mounting out on the feeder to supplement the reach of the substation circuit breaker. On long feeders, the substation breaker may only be able to detect a fault halfway to the end because the fault current magnitude near the end of the feeder becomes indistinguishable from the load currents at the substation. Reclosers would typically be located at 50 percent of the total feeder length to provide protection from that point to the end. Reclosers have instantaneous and time-delayed, element-tripping functions for interrupting fault current. They re-energize a distribution line by reclosing automatically after a period of de-energization (called dead time, typically from 12 cycles to 90 seconds, depending on the utility protection hardware and operational philosophy). The operation of reclosers out on feeders is coordinated with the operation of circuit breakers located at the substation and protects the entire feeder. The recloser must operate faster than the circuit breaker to selectively remove faulted segments of the feeder, and not the entire feeder. If a fault is permanent, the recloser locks open after a preset number of operations (usually three or four), isolating the faulted section from the main part of the system. Each tripping operation has a different time delay; usually the first trip is instantaneous and each successive trip is longer. Recloser controls provide the intelligence that enables a recloser to sense overcurrents, select timing operations, time the tripping and reclosing functions, and lock out if the fault is permanent. Hydraulic controls are integrated into the reclosers used in single-phase applications and smaller three-phase applications. Electronic controls are used for larger, three-phase applications and are housed in a cabinet separate from the recloser. Reclosers can operate on one or more timecurrent characteristic curves and are coordinated both with the fuses at coordinating points in the circuit and with the relays that control the circuit breaker at the substation.

3-6

The Distribution System

For instance, on overhead lateral lines, fuses are sized to operate on a time-delayed trip but not during an instantaneous trip, giving the temporary fault a chance to clear during the first breaker operation (fuse saving). Relays are the devices that control circuit-breaking elements such as breakers, reclosers, and sectionalizers. They close their contacts to actuate the circuit that causes the device controlled by the relay to operate. Many types of relays exist, with time-overcurrent relays being some of the simplest. Time-overcurrent devices can be set to operate in a predetermined amount of time or to operate according to the magnitude of the current in the line. The time delay on these devices must be coordinated with other protective devices on the circuit. Other common types of relays are directional relays, which prevent network current from energizing the high side of their transformers and feeder circuits, and differential relays, a measure that ensures that the amount of current entering the line is approximately equal to the amount of current leaving it. Directional relays are not typically used on radial distribution circuits but may be necessary once DR is employed. Differential relays are often employed for substation transformer protection. In general, for radial circuits, the current-tripping or operating thresholds of these protective devices decrease with increasing distance from the substation, in keeping with the expected behavior of the current in the lines. Engineers typically employ relays and a feeder breaker at the substation, fuses at laterals, and fuses at distribution transformer banks. In many cases involving longer feeders, a line recloser is added near the maximum reach of the substation breaker. In addition, many utilities use automatic sectionalizers to isolate faulted sections of distribution circuits. Automatic reclosing devices are often employed to restore power instantaneously after a temporary fault so that customers will experience only a momentary outage rather than a long-term outage for these faults. Fault protection devices are selectively coordinated so that only the device that is closest to a fault on the feeder will clear it. Such practices help minimize the number of customers who experience an outage from a fault and minimize damage to utility equipment. It must be noted that this coordination of the protective devices is made possible in the current radial systems by the assumptions noted above. Most importantly, that real power is flowing in only one direction and current is decreasing with increasing distance from the substation. Because the success of most of the protection practices discussed is based on radial power flow and no additional generation sources located on the circuit, the introduction of DR can interfere with the operation of these devices when the level of injected fault current from the DR becomes large enough to invalidate the basic assumptions. The details of some of these problems are discussed in Chapter 4.

3-7

The Distribution System

Networks
Low-voltage, network-type distribution systems, typically located in cities, serve only a small portion of the total circuit-miles of distribution within the United States. However, because of the high load density (and high customer density) of cities, they serve a large portion of the total customer load in this country! Networked systems are much more reliable than radial systems because they provide redundant paths for power to flow to the load. Loss of any one path does not result in loss of the ability to support the load. Networks are generally designed so that they can serve all loads with the loss of one primary feed to the low voltage network (This is called an N-1 design contingency). Some are even designed so that they can handle the loss of two primary feeds to the network (N-2). Examples of networks can be found in Manhattan, Boston, Atlanta, Dallas, Chicago, San Francisco, Seattle, and other major metropolitan areas. There are basically two types of networkthe grid network and the spot network (see Figure 3-4 and 3-5).

Figure 3-4 Low-Voltage Spot Network Configuration (Note: Arrows Show the Normal Direction of Power Flow.)

3-8

The Distribution System

Figure 3-5 Portion of a Low-Voltage Grid Network

Low-voltage, AC, secondary network systems supply combination lighting and power loads at nominal voltages of either 208 wye/120 volts or 480 wye/277 volts. Grid networks serve large areas of a city (10 by 10 city blocks, usually serving the smaller buildings of approximately 10 stories in height or less), while spot networks serve large, concentrated loads such as high-rise buildings. In a low-voltage, grid-networked system, the secondary mains that serve the end-use customers directly are connected together into a mesh or web structure. Power flows into the network from several primary feeders, where each feeder is connected through a transformer, a network protector, and a fuse.

3-9

The Distribution System

In the feeders, power is meant to flow in only one direction: toward the network formed by the secondary mains. Once in the low-voltage network grid, power can flow bi-directionally, but on the primary feeders, it is imperative to system protection that the power flow toward the lowvoltage network grid only, and not backwards from the network through a primary feeder. The advantage of the parallel feeder system is increased reliability; one feeder can be faulted or taken out of service for maintenance without an outage on the low-voltage grid. System protection, however, is more complex and expensive than it is in a radial distribution system. Network protectors are used to ensure the safety and reliability of the low-voltage grid. They prevent the back-feed of power through the network transformer into the primary feederthis is a necessary function if the primary feeder or network transformer becomes faulted. Installed between the network transformer and the secondary mains or network bus, the network protectors are essentially circuit breakers operating under the control of directional relays, and must trip under an extremely slight, reverse power flow (usually less than 0.5 percent of the rated forward flow). When a fault occurs on one of the primary feeders, reverse power flow can exist to the faulted feeder by way of the network transformer as well as fault current flow from the substation bus into the primary feeder. The primary feeder must be isolated from the source at both ends to prevent continued flow of fault current. The substation primary feeder breaker will trip at the substation end and the network protector, because of the reverse power flow, will trip at the network transformer end of the primary feederthis should fully isolate the faulted cable (see Figure 3-6). Because the reverse power trip threshold is small and is essentially instantaneous (occurs in just a few cycles) in most network protectors, DR has the potential to trip network protectors if it is interconnected on the low-voltage network and results in power export back to any of the primary cables. The details of this are discussed in Chapter 4.

3-10

The Distribution System

Figure 3-6 Example of Network Protector and Primary Feeder Circuit Breaker Isolating a Faulted Primary Cable (Note: Arrows Show Direction of Power Flow to Fault.)

Voltage Regulation of Power Distribution Systems


Maintaining proper voltage on the power system is critical to system performance, efficiency, safety, and customer satisfaction. Electric utilities use various combinations of equipment to control the voltage. These may include load tap changing transformers located in the substation, and feeder voltage regulators and switched capacitor banks (see Figure 3-7). Voltage delivered to customers is usually controlled within the bands specified by American National Standards Institute (ANSI) standard C84.1. Two levels existthe most stringent being the ANSI C84.1 Range-A requirement that says that the service voltage delivered to customers should be within 5% of nominal. The Range-B requirement allows slightly broader excursions for brief periods (+6% to 8%). Some utilities are required by local regulators to maintain voltage over a tighter range than ANSI limits. DR equipment can cause voltage excursions greater than or less than the required limits on radial systems by changing the direction and size of power flows downstream of voltage regulation 3-11

The Distribution System

equipment. DR may interact with any of the devices shown in Figure 3-7. Chapter 4 discusses in detail the impact of DR on voltage regulation.

Figure 3-7 Voltage Regulation Devices Used on Distribution Systems Include Load-Tap-Changing Transformers, Feeder Regulators, and Switched Capacitors

3-12

4
THE IMPACTS OF DISTRIBUTED RESOURCES ON DISTRIBUTION SYSTEMS

Introduction
The first step in assessing the impact of DR is to recognize that the distribution system sensitivity to distributed resources is determined not only by the size, location and type of DR, but also by the design characteristics of the distribution system. To evaluate the system impacts, engineers and planners need to collect data that can properly describe both the DR and the utility system to which it will be connected. Only then can the system impacts such as the influence on voltage regulation, fault levels, power quality, reliability, harmonics, stability and other performance characteristics be assessed. This section of the report focuses on the potential impacts of DR, identifying how they arise, how they can be predicted, and how they can be avoided if they are of a negative nature. It should be noted that many of the system impacts are discussed in the context of larger DR or higher aggregate amounts of DR placed on the system. For small generation systems and low aggregate amounts of DR, many of these impacts simply are not significant.

Data Needed to Evaluate DR Impacts


When any DR site is to be connected to the distribution system, some key issues and data to be considered include: Size rating of the proposed DR Type of DR power converter (static or rotating machine) Type of DR prime energy source (such as PV, wind, or fuel cell) Operating cycles (output level versus time of day, stops and starts per day) Fault current contribution of DR (How much? How long?) Harmonic output content of DR DR power factor under various operating conditions Location of DR on the distribution system Locations and settings of voltage-regulation equipment on distribution system Locations and settings of equipment for overcurrent protection on distribution system 4-1

The Impacts of Distributed Resources on Distribution Systems

Impedance of distribution system at point of interconnection Capacity of DR relative to feeder capacity at point of interconnection Locations and aggregate capacities of other DR DR output or rating relative to feeder load at point of interconnection The feeder grounding requirements and configuration The DR interface transformer configuration and grounding configuration DR protection functions and settings Other site-specific issues

The above list is generically representative of the information needed for a DR impact study. However, keep in mind that the level of detail and scope of data required depends on the context within which the DR is being applied. A single, 7-kW fuel cell applied at a residential customer does not require a comprehensive review of data for the entire distribution circuitrather it just needs limited information on the system in the local vicinity of the fuel cell site. On the other hand, a program to apply three hundred, 7-kW fuel cells to a feeder all at once would generate an aggregate capacity of 2.1 MW and would certainly need a comprehensive review of the feeder characteristics. As another example, a single, 3-MW, reciprocating-engine generator site would need information such as substation breaker trip settings and fault level impacts at all points on the feeder, because it is large enough to significantly impact the entire feeder. Engineers and planners working with DR need to always consider the context of their particular application and filter the data requirements down to the minimum needed to efficiently and safely interconnect to the system.

Sensitivity of Power System to Location of Distributed Resources


In real estate, the clich is that the three most important factors are location, location and location. For DR, this clich has much validity because the location of the DR on the distribution system is an extremely important factor in determining its impact on the system. The distribution system characteristics change dramatically from the substation bus to the customer-use level. Opportunities exist to connect DR anywhere on the system; including the distribution substation bus, the distribution primary feeder, a distribution secondary system, or imbedded within the power system at a customer site. The last location is probably the most likely for customer-owned DR at commercial or industrial facilities. However, independent power producers that own wind farms or other DR generation that intend only to sell to the utility company or bulk power market may connect at locations not imbedded within customer facilities. Utilities that operate their own DR for system support

4-2

The Impacts of Distributed Resources on Distribution Systems

reasons would often elect to connect at locations where a specific customer doesnt existsuch as at the substation or on the primary feeder with a portable support trailer (that is, a mobile DR). The farther the DR interconnection point is from the substation source, the weaker the distribution system to which it must interconnect. The term weaker is used to refer to the impedance of the system. A system with higher impedance is weaker than one with low impedance. System impedance determines the current the distribution system is capable of supplying with a given voltage drop. The higher the impedance the less current the system can supply. One of the best measures of the impedance that is quickly obtainable from any utility is the data provided in a fault (or short-circuit) study of a distribution circuit. These studies are routinely performed and show the short-circuit levels at all points on the system. Utilities use them to select the sizes of fuses, interrupting ratings, and trip settings for their protective devices. The short-circuit levels are directly related to the system impedance and can be used by DR planners to determine how weak or strong a given system is at a particular point. Short-circuit levels vary greatly at various points on a power system and depend on many factors including the substation transformer impedance and the distribution circuit conductor type, wire spacing and distance from the substation to the fault point. As the distance of the fault from the substation increases, the length of conductor through which fault current must pass increases, and hence, the impedance rises significantly. The type of fault, such as line-to-line or line-to-ground, will also determine the impedance seen. Typical fault currents at substations are anywhere from 3000 to 20,000 amperesdepending mainly on the size and characteristics of the substation transformer. Some substation fault levels are even higher if extremely large transformers are used. At 1 mile from the substation, the primary feeder fault currents decline to less than 6000 amperes, in most cases, and the conductor impedance becomes by far the dominant factor beyond that point. At the ends of extremely long, weak, rural feeders, primary feeder fault currents may possibly be less than 200 amperes. One way to estimate the impact of a DR on a utility system is to compare the ratio of rated output current of the DR to the available fault current of the utility system at the point of interconnection. As a general rule, DR units that have output currents greater than 1 percent of the available utility system fault current at the point of connection can cause noticeable impacts on voltage regulation, power quality, and voltage flicker. This is a general rule and the actual threshold where the impact becomes noticeable may be lower or higher depending on many factors. Based on the preceding discussion, it is clear that the substation with its higher fault current is usually able to absorb greater DR generation than a location at the end of a long, rural feeder that would have low fault currents. For example, using the general rule of thumb discussed above, a 13.2-kV substation with 7000 amperes of available fault current can handle 70 amperes of output from a DR before noticeable effects on voltage occur on the substation bus. 4-3

The Impacts of Distributed Resources on Distribution Systems

This is equivalent to a three-phase DR rating of approximately 1600 kVA. Whereas, at the end of the long, rural feeder with 200 amperes of fault current, the DR output limit is approximately 2 amperes; this translates into a three-phase, DR rating limit of approximately 46 kVA. These are, of course, greatly simplified examples meant for illustration purposes only. Many other factors, such as the need for voltage support and capacity support, will play a role in either increasing or decreasing these limits in a real world application. For example, when voltage regulators are used in line drop compensation mode, locations farther out on the feeder may actually be able to handle more DR than locations near the substation. It is also important to recognize that the limits just calculated considered only the distribution primary impedance and not the limits on the much weaker secondary system. If we start considering the secondary (low-voltage system) impacts, then significant impacts show up with even smaller DR. For example, DR smaller than 10 kW could cause noticeable voltage impacts on residential secondary circuits. Such impacts could affect not just the customer where the DR unit is installed but many other customers as well because most suburban distribution transformers serve anywhere from 2 to 8 residential customers. When applying smaller DR on such secondary circuits, the effects on the adjacent customers must be considered. Many commercial customers and most isolated rural customers wont use a shared distribution transformer on the secondary, and so, the secondary impacts on adjacent secondary customers can be ignored in those casesbut are still important for the actual customer applying the DR.

Distributed Resources at the Distribution Substation


The substation is the strongest point of the distribution system and can generally handle more DR capacity than other points on the system. For average-sized distribution substations with a typical transformer capacities of 12 to 20 MVA and typical impedance ranges of 8 to10 percent used in the transformer designs, DR capacities up to 2 MW can generally be accepted without much concern that voltage quality will be adversely impacted. Even larger DR can be successfully interconnected to the bus. For example, DR units equivalent to the substation transformer rating have been successfully operated. However, in these cases, much more care is required to assess interactions with power system equipment and often these larger installations require quite a bit of additional protection and control equipment to be installed at the substation and elsewhere. When DR is connected at the substation bus, the DR acts much like another transformer providing power to the distribution feeders. At this location, it is typically upstream of the various protective relay-sensing circuits for feeders and thus should have less influence on their protective relays. Of course, DR could still raise the fault levels and it still has the potential to backfeed into the transmission system and cause problems with the substation and transmission system protection. It also can interfere with the substation transformer voltage regulation controlsso all of these 4-4

The Impacts of Distributed Resources on Distribution Systems

issues must be evaluated if the DR is large enough. The addition of large DR at the distribution substation may also require possibly upgrading the following circuit-breaker ratings: Continuous, momentary, and interrupting current ratings Transient recovery voltage (TRV) Out-of-phase switching rating of circuit breakers Continuous current ratings of switches Protective relaying

Although locating DR at the substation bus avoids some of the issues that arise when DR is installed out on feeder lines, significant problems still exist that can arise if a large distributed generator is not properly integrated with the power system at this location. For example, significant interference can exist with the substations load-tap-changing transformer. To understand this, first consider that the voltage of distribution feeders emanating from the substation is normally regulated by a load-tap-changing (LTC) transformer. The LTC transformer boosts its output voltage according to the measured load at the transformer terminalsthe greater the load the more voltage that is applied. This is to compensate for voltage drop along the feeder lines that will occur as the load increases and helps to maintain proper voltage out at electric customers. If a large DR is installed in the substation so that it shields the current sensor from the true current on the line, then the LTC control will not boost the voltage sufficiently (see Figure 4-1). This will result in low voltage at the end of the regulation zone (see Figure 4-2). Of course, the DR would need to be fairly large at average size substations to shield the substation transformer from enough current to have a significant impact. A 500-kW DR on a typical, 12-MVA substation bus would not be a significant threat in this regardit is just too small. On the other hand, if generation size exceeds approximately 15 to 20 percent of the station loading, it can begin to cause problems in this respect. A DR as large as the load at the substation would almost certainly cause serious problems unless the appropriate provisions were made. Modifying the regulator controls so that they account for the DR current contribution or adjusting the regulator settings to disable the line-drop-compensation can avoid these problems. A DR at the substation that is large enough to cause serious voltage regulation issues, may also export power from the distribution system into the transmission system, could also potentially cause serious protection problems with the transmission interconnection. The protection, looking back into the transmission system, needs to be evaluated for these larger DR to ensure that there is no interference.

4-5

The Impacts of Distributed Resources on Distribution Systems

Figure 4-1 Placement of Distributed Resource at the Substation Will Generally Impact the Feeder Less Because of the Lower Impedance of the Substation

Figure 4-2 Voltage Profiles With and Without a Large Amount of Distributed Resources Connected at the Substation

4-6

The Impacts of Distributed Resources on Distribution Systems

Distributed Resources on the Primary Feeder and Secondary Lines


Distributed generation resources that are located on a primary feeder (or a secondary line) will have more influence on the distribution system operation than those of the same size load located at the substation because of the higher impedance of the distribution system at these locations. Interaction is also generally increased with the overcurrent protection system on the feeder because the DR can act as a source on the load side of protective devices and pass reverse power through them under some conditions. Because these are not generally configured to handle reverse power because of the radial protection philosophy of most distribution systems, this potentially can cause devices to malfunction, can damage equipment, and pose a safety hazard. Figure 4-3 shows how a feeder breaker could be nuisance tripped by such fault contributions. Compared with the substation bus, feeder locations are more sensitive to the DR contribution to fault currents because as a ratio of the utility system contribution to fault current level, the DR contribution will be higher for a given size DR. As a general rule, fuses, circuit breakers, reclosers, and other protective devices may begin to be interfered with if the DR changes the fault levels by more than 5 percent. The transformer winding arrangements associated with some generator installations can also alter fault levels even when the generator itself is not runningso these need to be assessed too!

Figure 4-3 Fault on an Adjacent Circuit Causes a Nuisance Trip of a Distribution Feeder because of the Backfeed of Fault Current from Distributed Resources (G1, G2, and G3)

4-7

The Impacts of Distributed Resources on Distribution Systems

Power systems vary so much that each must be considered, based on its own local characteristics, to determine the DR capacity thresholds where significant system impacts emerge. An extremely strong primary feeder location might be able to handle many megawatts of connected DR without incident. On the other hand, a weak primary feeder location may handle only a few hundred kilowatts before encountering difficulty. A secondary power system would generally handle much less power than the primary system but a wide-range of capacity capabilities possibly exist that depend on the secondary system design. For example, a residential service fed by a small transformer might handle only a few kilowatts, whereas a large commercial customer with a three-phase, 480-volt transformer may be able to handle many hundreds of kilowatts. Measures have been used in the industry to determine the appropriate size limits for DR that could be applied on the primary and secondary power system without serious adverse impacts on voltage and other parameters. Terms called stiffness ratios or stiffness factors have been recognized as one way to predict the impact of DR on the system and help determine if special requirements are needed. These ratios compare the strength of the utility system to the strength of the DR being connected. The term stiffness originates from long-used, electrical engineering terminology where a power source is considered stiff because no matter how much current you pull from the source, the voltage wont greatly changethe voltage is said to be stiff. One commonly used stiffness ratio compares the ratio of the available distribution system fault current at the point of interconnection to the steady state, full-load, DR-rated output current. Another one essentially compares the fault current of the power system to the fault current of the DR. Lets consider the first ratio as an example. In this case, a 100-kW, three-phase DR connected to a 13.2-kV primary will have a output current of 4.37 amperes per phase at its rated power level. If this DR was connected at a point on the distribution system with a fault level of 2000 amperes, then the stiffness ratio is 2000 divided by 4.37, which is 457.6. This is a high stiffness ratio that would predict little impact on the system. As a rule of thumb, applications with stiffness ratios that are greater than 100 are not likely to create voltage problems. Note that we calculated the stiffness ratio for the primary. If we had done it at the secondary voltage level, then the answer would be extremely different. The secondary fault current might be 15,000 amperes at a small, commercial service entrance (for example a fast food restaurant), and the output current of the generator at the secondary voltage level of 208 volts would be 277 amperes per phase. Here the stiffness ratio is only 54. Here you can see the difference in the capacity of the primary distribution system compared to the secondary system. When do you use the primary stiffness ratio or secondary system stiffness ratio? The answer depends on the intended objective of the analysis. If the objective is to identify potential voltage

4-8

The Impacts of Distributed Resources on Distribution Systems

and power quality problems at the DR site itself, then you would use the stiffness ratio on the secondary right at the DR connection point. If the objective is to identify potential problems for the utility company system and adjacent customers, then the location where the nearest adjacent customer is connected to the power system would be a good spot to determine stiffness factor. This location might be on the secondary or the primary depending on the nature of the distributions at the site of connection. At many sites, the customer is served with a dedicated distribution transformer and the secondary is not shared with another customer. Thus, in these cases the stiffness ratio on the primary side of the transformer can be used if the impacts on the DR site are of no concern (see Figure 4-4, Case B). On systems where the secondary serves several customers, the stiffness ratio can be determined at the point shown in Figure 4-4, Case A.

Figure 4-4 Evaluation Points Where the Stiffness Ratio Should Be Evaluated

4-9

The Impacts of Distributed Resources on Distribution Systems

In the preceding discussion, a stiffness ratio of 100 or higher was suggested as a safe value for DR. This is just suggested as a conservative rule of thumb. It is certainly possible to have successful DR installation with stiffness ratios lower than 100. It is just that serious problems are more likely to be created and issues for the system may require additional protection and control equipment to help mitigate negative impacts. In most cases, the DR owner/operator would be expected to pay for the cost of such modifications. For large DR of several megawatts or more on typical distribution feeders, no location on the distribution feeder would have a stiffness ratio greater than 100. The term stiffness ratio has also been used by The Institute of Electrical and Electronics Engineers (IEEE) in the development of its draft standard IEEE P1547. The IEEE P1547-D8 stiffness ratio is defined as the ratio of the electric power system supplied fault current (including the DR contribution) to the fault current of the DR alone. This approach is a different way of calculating essentially the same thingthe relative size of the DR compared to the capacity of the power system. The IEEE ratio is useful for evaluating the impact of DR on fault levels. Both ratios discussed so far, when used together, can describe the impacts on fault levels and voltage regulation and serve as helpful tools to quickly screen DR applications for the severity of their potential impacts. Stiffness ratio serves as one tool among the many needed to successfully screen DR applications for their impact on the system. So far in discussing stiffness ratio, we have only dealt with a single DR unit. However, consideration of the total aggregate capacity of DR connected to the distribution system is also extremely important. As an example, a single 10-kW photovoltaic DR on a typical distribution feeder serving 2000 homes has insignificant impact on power flows on the feeder primary. However, if 500 homes on that feeder install 10-kW solar arrays, then the aggregate output of these units is 5 MW on a sunny day! This would have a major impact on the feeder power flow and could actually export power back to the substation bus. In introducing DR to the system, we need to consider the aggregate capacity of generation. If planners are evaluating the long-term impacts of DR then, in some cases, aggregate stiffness ratios may be used for screening. Technical limitations exist to this approach, however, and more traditional and time-consuming short-circuit studies may be needed to fully evaluate the impacts of multiple DR feeding into the system. There are technical questions that the power industry is also still trying to resolve such as should aggregate DR capacity be treated the same way as a single large DR unit, and how will all the different units interact under steady state and dynamic conditions? Until these questions are resolved with experience, the industry should be conservative in establishing limits for the aggregate capacity that is allowed on systems. Many engineers dont feel comfortable with more than 10 percent aggregate capacity on the system.

4-10

The Impacts of Distributed Resources on Distribution Systems

Losses and Reactive Power Flow


The U.S. T&D system is estimated to have 8-10 percent total losses (it is approximately 9092 percent efficient). Ideal placement and dispatch of DR can reduce some of these losses. A DR will reduce load currentrelated losses on any section of the line or at any transformer where it reduces the current. Ideally, to get the maximum reduction in losses, the distributed generator must be located at the correct point on the feeder, dispatched around the time of peak system losses and operated at the optimal power output level. A DR that is too large for a given feeder location (meaning that it exports too much power), could actually increase T&D system losses. Customer-owned DR that does not export power but instead acts to reduce the load at a specific customer site will always reduce lossesbut the greatest benefits to the distribution system occur for non-exporting DR at the end of radial circuits operated at the time of peak system load. Placement of DR to optimize losses on a power system (obviously more of concern for utility-owned DR than customer-owned DR) is similar to placement of capacitors for the same purpose. The main difference is that capacitors can affect only reactive power flow, whereas, DR can impact both real power and reactive flows. The ideal feeder location for a single, DR site to minimize real power related losses is at the point on the line where the output of the DR is equivalent to twice the loading at that point. Thus, half the power flows to the downstream load and half flows upstream (see Figure 4-5)this minimizes the transit distance of power from sources to loads and therefore minimizes losses.

Figure 4-5 Optimal Location for a Distributed Resource to Reduce Feeder Losses

4-11

The Impacts of Distributed Resources on Distribution Systems

Impact of Power Factor on Losses


Any DR operating at a lagging power factor condition is increasing reactive power demand on the feeder. For example, an induction generator operating at a power factor of 0.8 will inject real power but will require (draw) some significant reactive power for excitation. Thus, if sited as shown in Figure 4-5, it will help reduce losses related to real power flow but will actually increase losses because of the reactive part of its power flow. Overall, the loss reduction could still be positive at rated output of the generator because real power losses would normally decrease more than reactive losses would increase. At much less than rated output (for example, less than 25 percent), the power factor declines substantially and some types of generators could actually result in a net increase in losses because their reactive current requirements are greater than the real current they inject! To avoid this, DR units should be designed to operate at fairly high power factor over most of their expected power-output operating band. Most inverter technologies in the U.S.today and all synchronous generators can be controlled to operate extremely close to unity power factor near their rated load. Induction generators would normally operate at a power factor of 0.8 or less if not compensated with capacitors. Induction generators can be compensated to a power factor near unity with capacitorsbut this must be done carefully (with the right controls) to avoid self-excitation problems that could lead to unintentional islanding problems. Its also important to recognize that, in addition to reducing losses, operating DR so that it is not a reactive burden on the system will help reduce the use of T&D capacity to support reactive power requirements. In general, having good power factor of connected generation helps the utility system operate efficiently and avoids unneeded infrastructure costs. As a result, many utilities have guidelines for the power factor of connected generation that specify that it must be at least 0.9 or even 0.95 or better. IEEE standards such as 929-2000 and the under-development P1547, also specify power factor requirements.

Current Ratings and Capacity Issues


With the addition of large quantities of distributed generation to the distribution system, the continuous current rating of circuit breakers and switches may have to be increased because of increased current flow. This would be the case where the DR capacity is so large that it fundamentally changes the flow patterns in a major way on the distribution circuit. This type of effect would tend to occur only with a large DR of many megawatts capacity or large, aggregate DR capacity. In the case of any single smaller DR or where the aggregate capacity is less than 10 percent of the feeder peak demand, it would be unlikely that any changes in ratings would be required from a continuous-duty perspective. The momentary and interrupting current ratings of existing and planned, new circuit breakers should be evaluated when DR is expected to have a major impact on fault levels. The fault characteristics of the DR need to be understood so that ratings of all switchgear can be 4-12

The Impacts of Distributed Resources on Distribution Systems

coordinated with the DR. The condition that poses the highest possible fault current levels should be evaluated. An interlock system can be used to prevent switching operations that can cause conditions of excessive current. ANSI/IEEE C37 contains necessary information to rate both high- and lowvoltage breakers. Again, in most instances of low-DR penetration, this would not be an issue on the distribution system primary. However, with the addition of a larger DR and/or higher aggregate capacity levels of DR, existing switchgear would generally need to be evaluated to see that ratings were not exceeded. On the secondary, the addition of a moderate amount of DR capacity could exceed the ratings of some switchgear. Commercial and industrial customers installing DR should assess their switchgear to ensure that ratings are not exceeded. DR will also influence the currents on the feeder conductors. Suitable capacity must exist to handle any increased current flows introduced by DR. Capacity limits on distribution systems are typically determined by either thermal or voltage-drop considerations. A distribution system is said to be thermally limited when, at high loading levels, the wires or equipment on the feeder reach their maximum operating temperature before any voltage problems develop (a voltage problem is an out-of-range voltage). A system is voltage-drop limited when out-of-range voltage (according to standards such as ANSI C84.1) occurs at a loading level that is less than the thermal loading limit. Rural radial distribution systems tend to be voltage-drop limited and shorter, urban/suburban, radial distribution systems tend to be thermally limited. Low-voltage network systems are also thermally limited. DR must be coordinated with the system thermal and/or voltage limits where it is applied. Too much DR at any point on the system can cause thermal overload or out-of-range voltages. DR can also affect the stability of the power system. However, in most current applications, stability is less of an issue than the other factors we have discussed. It will become a significant issue in the future if penetration of DR increases and DR becomes a key resource for supporting the bulk power system and the distribution system. Later in this report, a detailed discussion of stability issues can be found. DR capacity limits must also consider emergency and contingency conditions on the distribution system. Feeders may be reconfigured during upstream faults to allow backfeed from another feeder or they may be required to pick up the load of another feeder. The capacities and placement of DR must consider these possibilities. If DR is relied upon as a resource (to defer T&D investment), then the design of the distribution system needs to consider emergency and normal conditions where the DR is not available. Conventional distribution generation technologies such as reciprocating engines and combustion turbines may have availabilities in the range of 0.90 to 0.98from a few days up to one month of down time per year! The feeder, if it is relying on this DR for support, should be designed to operate satisfactorily during the period when the DR is unavailable. Planned DR outages for maintenance should be scheduled for off-peak weeks (such as spring or fall), if the DR is intended for feeder support to reduce the probability of a capacity shortfall. 4-13

The Impacts of Distributed Resources on Distribution Systems

Safety
As DR is introduced to the system, it creates a completely new concern with regard to safety since utilities no longer have complete control over the energizing of the feeder. Modern distribution system design, protection hardware and practices, and personnel safety hardware and procedures have evolved based on a relatively small number of centralized generation units delivering power over radial feeders. The operational status of these systems is easily known and understood by utility personnel. Electric utilities have work procedures for maintenance of both energized and de-energized parts of the distribution system. In general, the safety procedures for de-energizing an electric distribution system include six steps: notification, identification, switching, tagging, testing, and grounding. In the last step, all potential feed points to a work area are grounded and other energized equipment nearby is protected with insulating shields and covers. Under normal conditions, when the utility disconnects from the feeder so too must DR because it cannot support the load. This process is normally completed in a matter of cycles. Under certain conditions, however, where the size of the DR approaches that of the system to which it is connected, islanding may occur. An obvious concern is that linemen may be exposed to the unexpected startup of a DR unit. DR must have controls and provisions that prevent islanding. Protection circuits are required that can sense that a generator is islanded and quickly isolate it from the system. In addition, a means of permanent disconnect (such as a visible break, utility accessible, lockable, load-break-rated disconnect switch) is required as further protection by most utilities. Utilities must review and possibly revise de-energizing procedures in the increasingly likely event that DR is present on a feeder. If utility procedures provide protection from contact with other circuits downstream from the work location, they probably also provide protection from backfeed from a DR installation. Any revised procedures must not only comply with existing Occupational Safety and Health Administration (OSHA) regulations concerning disconnection and tagging of sources that could feed into the fault location, but must also provide workers with a maximum safety margin. Expensive, live-line techniques may be required if this is not possible. The topic of islanding and de-energizing islands that develop is discussed later in this chapter. It is important to recognize that DR that ordinarily are not intended to export power into the distribution system still have the potential to do so if the relaying and protection is incorrectly applied or fails. No relaying scheme is 100 percent reliable, and anti-island protection can fail under conditions where a balanced, load-generation island forms. Also, even with protective relays and anti-island protection working properly to prevent export of power, transient conditions will exist wherein, power will be briefly exported for a period of a few cycles up to several seconds (the period of time required for the protection to respond). Under these conditions, the role of the DR in safety and system protection issues cannot be ignored. 4-14

The Impacts of Distributed Resources on Distribution Systems

Impact on Fault Conditions


DR can change the dynamics of the distribution system during faulted circuit conditions. Under such conditions, some key effects include (1) changing the magnitudes of fault currents, (2) changing the direction of power flow, (3) causing a mismatch between the actual fault current versus what is measured by utility company protective devices, and (4) causing potentially higher transient recovery voltages on switchgear. These effects can potentially interfere with the operation of fuses, circuit breakers, reclosers and other distribution system protection devices and may lead to reliability and safety problems if not appropriately addressed. The various fault condition impacts are discussed in detail below. Transient Recovery Voltage Transient recovery voltage (TRV) is the voltage that is created on a distribution circuit during the fault current interruption process of a fuse or circuit breaker. At time zero, when current has just been interrupted (for instance a circuit breaker opening to clear a fault), the system voltage oscillates in accordance with its natural frequency. The TRV is impressed across the opening breaker contacts and stresses the gap insulation. The interrupting media, following current extinction, attempt to go from a state of good conduction to one of a good insulation. The interrupting media must be a sufficient insulator to prevent current flow across the gap because of the TRV. If the insulation is able to prevent arcing across the TRV, then a successful interruption occurs. If not, the arc is re-established and the interruption process is attempted again. This process continues until either successful interruption or breaker failure occurs. Adding DR or a currentlimiting reactor1 may result in more severe transient recovery voltage. If the TRV imposed by the DR is large enough, then the breakers may need to be upgraded to accommodate it. Some interrupting devices have not been rated for or designed to interrupt the voltage between two energized regions of the systemthat is the utility system and an island2 created with DR. Fault Levels The addition of distributed generating devices on the distribution system will add to the fault current levels on the power system (see Figure 4-6). Whether these currents seriously impact the system or not depends on many factors including the amount of fault current contribution by the DR relative to the pre-existing utility system fault levels, the location of the fault, directional characteristics of the contributing current(s), and the types of protection schemes employed by both the utility system and the DR.

A reactor (inductive coil wound on an iron core) can be used to insert impedance into the power system to help limit the magnitude of fault currents. Some substations employ reactors to help reduce fault currents. 2 Note that regardless of the anti-island protection of a rotating generator, that an island always forms briefly for a few cycles until the DR is tripped offline by its protection.

4-15

The Impacts of Distributed Resources on Distribution Systems

Figure 4-6 An Example of Fault Contributions from Distributed Resources Being Added to the Contribution from the Substation Source and Increasing the Fault Level on the Lateral

The fault contribution magnitudes of generators are determined by the type of power converter (rotating or static) employed, the excitation system used on the generator, and the impedances in the intervening path to the fault location (this includes the step-up transformer and lines). Worst-case, peak, fault-contribution levels associated with an individual rotating machine (both induction and synchronous) can be quickly estimated by using a circuit model that considers the machines prefault output voltage divided by its sub-transient reactance3. Using this approach, it can be shown that for most synchronous or induction generators, a threephase fault at the terminals can produce an initial short-circuit current of 5 to 10 times the rated current of the machine for a few cycles following fault initiation (this is called the sub-transient period). After this period, comes the transient period where currents are two to five times the rated current for another 10 cycles or so. For solidly grounded rotating machines, the singlephase fault contributions are typically 30 percent higher than three-phase because of the nature of machine design (the impedance is effectively lower for this fault mode). Single-phase faults often pose a threat to the rotating machine itself because single-phase fault currents cause higher mechanical stresses on the generator windings than three-phase faults. Rotating machines often have grounding impedances intentionally inserted into the ground path,
Generator impedance can be characterized by three different impedance measures; the subtransient, transient and synchronous reactances. These impedances determine the response of the generator to a load step or short circuit during various time periods. When a fault occurs across the machine terminals, the current can be initially calculated by dividing the voltage of the generator by its subtransient reactance. Sub-transient reactances are the lowest of the three reactances and are important during the first 1-5 cycles of an event and thus dictate the worst case fault contribution from a machine.
3

4-16

The Impacts of Distributed Resources on Distribution Systems

are ungrounded, or use transformer arrangements that help limit their exposure to single-phase fault contributions so that the machine is not damaged. The fault current contributions from induction and synchronous machines decay to essentially zero amperes within approximately 10 cycles if no excitation is present. However, excitation of the induction generator may continue if any phase still has partial or full voltage. For example, voltage can persist on the unfaulted phases of three-phase systems when a single line to ground fault occurs. With continued excitation, the fault contributions will continue until the unit is tripped offline. For a synchronous machine excited by a source that can continue to supply exciter power during the fault, fault contributions will continue until the machine is tripped offline, even if the voltage on all three phases collapses to a level near zero. The amount of current potentially contributed by machines after the sub-transient and transient periods is dependent on the steady state impedance characteristics of the generator and the level of excitation that is sustained. This may be anywhere from a fraction of the rated current of the machine up to several times its rated current depending on these factors. The full-time frame of fault current contributions of a generator from the sub-transient period to steady state is shown in Figure 4-7.

Figure 4-7 Contribution of Fault Current From Typical Rotating Generators During the Sub-Transient, Transient, and Steady State Time Frames

Many utility system fault protection devices are impacted in the time frame of just a few cycles, so the transient and sub-transient periods of the fault contributions are important in the analysis of the protection impacts. While fault contributions could in theory last for many seconds in the steady state region of Figure 4-6, it is important to recognize that protection standards such as draft standard IEEE P1547-D7 and common, utility-specified, DR interface practices require that 4-17

The Impacts of Distributed Resources on Distribution Systems

generators trip off line upon detection of a fault. For faults on the utility system near the generator this may occur within 10 cycles or less, thereby limiting the fault contributions to no longer than the transient period. Static power converters (inverters) are another source of fault contributions, but they behave differently than rotating machines. The magnitude and duration of the fault current contribution from an inverter is dependent on the inverter programming. Some inverters have thermal protection limits that quickly (within a few milliseconds) limit the current output or shut the inverter off to avoid damaging transistors when faults are detected, and others can contribute many times the rated current of the inverter for many cycles or longer. The voltage conditions that exist at the inverter terminals during the fault will also determine whether the unit continues to contribute or if it shuts down extremely quickly. For utility system interactive inverter designs, if the voltage drops significantly or the phase angle changes too quickly across the inverter terminals, the inverter will typically shutdown within milliseconds. In such cases, the fault contributions are of negligible duration. For conditions where the voltage at the inverter terminals does not drop that much (perhaps because of a distant fault), the inverter may help feed the fault for several seconds. Most inverters will contribute much less fault current than a similarly sized rotating machine so inverters have less impact on system protection than rotating machines. Despite this, no general way exists to describe how an inverter responds to a system fault and what its contribution will beeach inverter design must be assessed individually against the standards it is built to satisfy and/or against test results on the unit. The possibilities range from essentially no fault contribution (only an extremely short 1-2 millisecond current pulse) up to 10 cycles or more of four times the rated current being supplied until the protection trips the unit off. An inverter that contributes current for only a few milliseconds is contributing current for less than 1/6th of a cycle and will have essentially no impact on the utility protection equipment regardless of its size. On the other hand, an inverter that supplies four times rated current for 10 cycles is contributing almost as much current as a rotating machine and will have a significant impact if it is large enough. If the fault contribution level of the inverter is unknown and its impact needs to be assessed, then use of the four-times-rated-output rule is a conservative approach that can be applied to determine the worst-case fault contribution. As an example of the contributions to fault current that can occur with various machines, lets first consider a typical 480-volt, three-phase synchronous, rotating generator with a power rating of 100 kW. This type of generator would contribute during its sub-transient reactance period approximately 500-1000 amperes of fault current for a fault at its terminals. On the primary side of the transformer (13.2 kV phase-to-phase), the current contribution would be only 18 amperes, which is generally not enough on a typical feeder primary to cause problems for any protective devices (there are some exceptions). On the other hand, a 2-MW-rated, synchronous generator could contribute 360 amperes on the primary and could significantly change the fault level on the distribution systemespecially at weaker points on the distribution system. This could impact fuses, circuit breakers and other 4-18

The Impacts of Distributed Resources on Distribution Systems

devices (see upcoming sections on fuse saving, nuisance tripping, and other similar topics). So you can see that a single, moderate-sized system is not much threat, but a larger, 2-MW system can pose a threat. As a second example, lets consider an inverter and lets consider the worst-case scenario. A 10-kW, single-phase inverter operating at 240 volts could, in the worst case, contribute approximately four times its rated current or 200 amperes at its terminals. On the high-voltage primary of a 13.2-kV line (after the step-up transformer), this contribution would be only 1.3 amperes. This is small relative to the ratings of fuses used in the distribution transformer and is insignificant relative to the existing utility system fault levels. It would take many 10-kW units in aggregation to create the potential to significantly change fault currents. As an example, if two hundred 10-kW inverters were connected in aggregation, then the total contributions would be approximately 260 amperes on the primary. This level is large enough to significantly change fault levels on the distribution system. It should be noted that we have used the worst-case scenario in this discussionthat is a fault contribution of four times the rated power of the inverter. In many scenarios, the inverter would shut down in a few milliseconds and the contributions would not be significant even with hundreds of inverters! The preceding examples of fault current contributions from both rotating machines and inverters did not consider the distribution system impedance or utility source residual voltage effects. These factors will reduce the fault contributions of DR somewhat. The degree to which this occurs depends on the location of the fault and nature of the utility system design/equipment.

Impact of Fault Contributions of Distributed Resources on Lateral Fusing Practice


Utilities usually use fuses to protect single-phase laterals and occasionally to protect three-phase branches from faults and overloads. Common fusing approaches assume that many faults on radial overhead distribution circuits are temporary in nature, which means they can be cleared by de-energizing the line for a few seconds or less. Use of a fuse to clear a temporary fault is often avoided because it requires time to replace the blown fuse and the section of line impacted would experience a sustained, long-duration interruption. Instead, circuit breakers or reclosers upstream of the fuse are set to trip before the fuse blows. Then, once the fault is cleared, the breaker recloses a moment later and the feeder is back to normal. This practice is called fuse saving and it helps improve the reliability of the power system. For fuse saving to work, the breaker must interrupt the fault flow extremely fast because fuses melt quickly. Without DR on the system, fuse saving requires careful attention to fault levels and the types and sizes of fuses and breaker settings used. The breaker must be tripped not only before the fuse melts but also before it is damaged. Sometimes the practice cant work because the fault currents are too high and the fuses will be damaged before the breakers can clear the fault current. With DR, fuse saving can be even more difficult to achieve because the fuse will see increased current and is damaged even faster. 4-19

The Impacts of Distributed Resources on Distribution Systems

Figure 4-8 shows an example of how fuse-saving coordination is impacted by DR. In this example, the fault level is shown before the addition of DR where the breaker can trip (clear the fault) before the fuse becomes damaged. After DR is added, the increase in fault current results in fuse damage before the circuit breaker trips. Note also that the fuse will see greater current than the circuit breaker because the DR fault current contribution impacts the fuse and not the breaker (see inset picture in Figure 4-8). This can further exacerbate the problemespecially if a time delay relay is used for the circuit breaker. Fortunately, a few smaller DR do not contribute enough fault current to degrade most fuse saving schemes. This would be a problem with larger DR or large aggregate amounts of DR.

Figure 4-8 An Example of How Fuse-Saving Coordination Is Impacted by Use of Distributed Resources

Fuse saving is only one aspect of fuse applications on the power system. Fault contributions of DR can also interfere with other fuse application issues such as fuse-fuse coordination, where several types of fuses are cascaded in series or with distribution transformer fusing. In the case of either type of application, the added fault current could cause fuses to operate (melt) when they should notleading to decreased power system reliability. 4-20

The Impacts of Distributed Resources on Distribution Systems

As was the case with fuse saving, it would normally be only the larger DR or large aggregate amounts of DR that could cause this problem to occur. Careful attention to the fault contributions of connected DR can ensure that most of these problems are avoided or recognized before they become an issue. If a problem is identified, it can be corrected by adjustments to the fuse sizes, relay settings and/or DR tripping settings.

Impact of Reverse Flow Fault Contributions on Sectionalizers, Reclosers, and Circuit Breakers
Overcurrent protection devices such as circuit breakers, reclosers and sectionalizers normally operate by allowing currents below their tripping threshold to pass and will trip if currents exceed the trip thresholdfor radial distribution systems these devices normally sense magnitude and not direction of flow. Several devices can be cascaded in series at various points on the feeder and the trip thresholds and time delays are coordinated to ensure that the appropriate overcurrent device is used to isolate the fault from the system. This helps to minimize the number of customers interrupted during the fault. Sensing only the magnitude of fault current works fine for radial systems where the power only flows from the substation to the fault. However, this approach can be confused by the introduction of reverse power flow during faults when DR is present on the system. When a fault occurs upstream of a protective device such as a recloser or sectionalizer, the current contribution from any downstream DR passing through that device can fool it into believing that a fault has occurred downstream of the device. Depending on how the sectionalizer or recloser is programmed to operate, this can prevent the proper execution of automatic sectionalizing or auto-loop schemes that are intended to minimize the number of customers affected with an outage during a system fault. The example of Figure 4-9 shows the potential DR impact on auto-sectionalizing switches in an auto-loop scheme. First, lets consider the normally operating scheme without any DR present. In the normally operating scheme, if a fault occurs on Feeder #2 between the substation and the switch labeled C, then Feeder #2 is de-energized by the opening of the Feeder #2 circuit breaker. Then, seconds later, the auto-sectionalizing Switch B will open while the feeder is de-energized because its programming has been set to open on loss of potential (these switches arent rated to clear faults so they open after the feeder is de-energized). Switch C, that feeds the branch circuit, stays closed because it is programmed only to open if it had sensed fault current during the fault eventbecause the fault was upstream, none would have passed through it, so it stays closed. After a period of a few seconds, the auto-loop recloser, that has been normally open, senses the loss of voltage on Feeder #2 and closes into the backend of Feeder #2providing power to all customers on the back half of Feeder #2 while the front half is being repaired. This would also energize the customers on the branch circuit too. This whole process takes a few seconds and the customers on the back half of the feeder see a short interruption, while those on the front, see many hours while the fault is fixed. 4-21

The Impacts of Distributed Resources on Distribution Systems

Now lets look at what happens if a large DR is added to the circuit. DR is added on the branch circuit shown in Figure 4-9, then the fault contributions passing through Switch C could make it appear that a fault exists on that branch. The end result is that Switch C may open and prevent the backfeed of that section when the recloser closesthis results in a larger area (group of customers) without power during the fault and reduced the reliability of the power system. This example is just one of many possible complications that can arise with auto-loop and sectionalizing switch schemes. Fortunately, this type of problem would most likely arise only with fairly large DR.

Figure 4-9 Distributed Resource Located on the Branch Circuit Contributes Fault Current Causing Confusion to Sectionalizing Switch C

Another problem that can occur is that an adjacent feeder fault can cause a nuisance trip of a feeder circuit breaker because of the reverse fault current contribution from DR. As is shown in Figure 4-10, when a fault occurs on an adjacent feeder, generators G1, G2, and G3 all contribute fault current to the adjacent feeder fault. Fault contributions from these generators, if large enough, could make it appear that the feeder is faulted, so both breakers will trip. As stated before, substation feeder relays that control the circuit breakers do not normally employ reverse power flow detection and therefore can be fooled by these contributions to the adjacent feeder fault. The end result of such an event is to cause an unnecessary outage on an entire feeder circuit. Many of the problems associated with reverse flows because of DR can be mitigated by the application of directional relaying in substations and at other equipment where possible. However, this is not normal practice and could cost many tens of thousands of dollars in 4-22

The Impacts of Distributed Resources on Distribution Systems

engineering, hardware, and installation costs to upgrade feeder equipment. Some reclosers and sectionalizer switches are not easily upgradeable and fuses are not upgradeable.

Figure 4-10 Example of Nuisance Tripping on the Feeder With Distributed Resource

Size Limits Where Distributed Resource Fault Contributions Become an Issue


The fault current contributions needed to create the protection problems discussed in the preceding sections depend on the type of utility system device, its trip setting, intended application, and its location on the feeder. For example, the instantaneous-trip threshold employed by most distribution feeder circuit breakers is 1000 amperes or more. To exceed this threshold on a 13.2 kV system requires an aggregate generator capacity of at least 5 MW. For 4.8 kV feeder systems, a generator capacity of about 1.8 MW is needed to create about 1000 amperes of fault contribution. Out on the feeder, the protective device ratings and trip setting become smaller. On long weak rural feeders, reclosers, sectionalizers and lateral fuses may be activated by less than one hundred amperes of fault contribution. In these cases, DR of several hundred kilowatts rating could be enough to cause nuisance trips, miscoordinations and other problems. In the case of fuses protecting distribution transformers, most fuses of this type are rated anywhere from a few amperes up to a few tens of amperes. The fuse size depends on the transformer size and fusing practice employed by the utility company. Examining the melt-time curves for such fuses, it is clear that DR fault contributions should not ordinarily cause a distribution transformer fuse to melt. Some situations can be identified where there may be a small risk. An example is a 25-kVA transformer that serves four houses. 4-23

The Impacts of Distributed Resources on Distribution Systems

If each house is equipped with a 10-kW DR system, the total capacity connected is 40 kW. If the transformer is fused with a small fuse (such as 3-ampere rating), the fault current contribution could melt such a fuse under worst-case conditions. However, in most cases, the fuse would be larger than 3-amperes (more like 6 to 15 amperes) and this risk can be mitigated by simply not allowing the connected DR capacity to exceed the transformer rating. Overall, many factors determine the threshold size where DR fault contributions become large enough to significantly impact power system protection. These factors include the magnitude and duration of the current contributed, the location where the DR is interconnected, the type of upstream protective devices and their settings, and the complex interactions of various aggregations of DR on the system. As a general rule, if fault currents are not increased by more than 5 percent, then the likelihood of an issue is generally small, but there are some exceptions to thissuch as with network protectors. Note that the impact on power system fault current is not just determined by the DR but also by the type of grounding and transformer interface employed on the system. These grounding impacts are discussed in the next section.

Grounding Compatibility
In general terms, grounding refers to the method in which the power system is electrically connected to the earth. Some types of power systems have no direct connection of any currentcarrying wire to the earth and others have numerous connections. The configuration for grounding the power system is a crucial design characteristic of every power distribution systemimpacting its safety, protection, and how loads and generators can be connected to it. DR must be configured and operated in a manner that is compatible with the grounding design of the distribution system. Failure to achieve compatibility can lead to serious safety problems or damage to customer loads and/or utility company equipment. The specific meaning of grounding depends on the context in which it is used. It can refer to the grounding of the secondary system, the primary distribution system or just an individual generator or load. When utility distribution system engineers worry about grounding of the generator in relationship to the power system, they are usually referring to both the primary and the secondary power system. They want to make sure that the generator grounding should be compatible with both of those systems. Grounding compatibility can seem extremely complex to persons unfamiliar with three-phase power systems because engineers often discuss it using a vector analysis approach called sequence components or symmetrical components. These can be extremely complex for the uninitiated. So, for this discussion well avoid referring to sequence components as much as possible so as to make it palatable for those without extensive background in that analysis. In the United States, the most common type of primary power distribution system is the four-wire, multi-grounded-neutral system (see Figure 4-11). This system uses three phases (A, B, and C) and a fourth wire that is called the neutral. All four wires are used for three phase 4-24

The Impacts of Distributed Resources on Distribution Systems

sections of the line and any one phase and the neutral are used for single-phase sections. The neutral is grounded (that is, connected to earth) at various points. This earth connection is usually at all transformers, lightning arresters, and other equipment as well as at some poles without any equipment. On a typical distribution feeder of this type, anywhere between 100-1000 grounding (earth) connections could be found. The connection to ground is by rods driven into the earth (usually at poles), by connection to both a rod and a copper water pipe system (usually at customers), and to a large grid of buried wires and driven rods (at a substation). A key design feature of a four-wire system is that most of the distribution transformers are phase-to-neutral connected (note the houses served by the phase-to-neutral connected transformer in Figure 4-11.)

Figure 4-11 Grounding of a Four-Wire, Multi-Grounded-Neutral Distribution System

A much less common, but still widely used, grounding scheme in the United States is the ungrounded, three-wire distribution system. This scheme uses three phases (A, B, and C) but no neutral wire and so there is no ground connection of any current carrying wire. Figure 4-12 shows a typical, three-wire, ungrounded primary system. In this system all distribution transformers are phase-to-phase connected to the primary.

4-25

The Impacts of Distributed Resources on Distribution Systems

Figure 4-12 Example of the Three-Wire, Ungrounded Primary System

Regardless of the type of grounding on the primary distribution system, the secondary4 system may be configured as either a grounded or ungrounded systemthe secondary grounding employed depends on the nature of the secondary system design and the loads at the point of use. In most designs, the secondary is configured as a four-wire, grounded system. However, some industrial and commercial loads are connected as a three-wire, ungrounded system. Proper grounding analysis of DR will ensure compatibility with grounding for both the primary and secondary power system. This analysis must consider (1) the generator-winding configuration (or inverter arrangement), (2) its grounding point, (3) the interface transformer configuration, and (4) grounding of both the primary and secondary power systems to which the DR is connected. It is important to recognize that a grounding arrangement that is compatible with the secondary system may not be compatible with the primary system! Certain DR interface designs can be compatible with both systems and utilities and DR installers should aim for these designs. Good grounding analysis must also recognize that more than one transformer may be found between the DR output terminals and the primary distribution system.

The terms secondary and primary are being used in the context commonly employed by utility engineers. The secondary is the low voltage side of the power system, that is 600 volts or less, that is connected to residential and commercial loads most houses are served by a secondary voltage of 120/240 volts. The primary is the high voltage side of the distribution transformer; the feeder and laterals that operate at several thousand volts (e.g. 13.2, 12.47 or 4.8 kV).

4-26

The Impacts of Distributed Resources on Distribution Systems

For those readers not familiar with the engineering details of power system design, the role of the transformer in grounding is not intuitively obvious. This role is discussed in greater detail later, but for now it is sufficient to state that some transformer types do not allow ground currents to pass through then. This means that a generator can have proper grounding on the secondary side of the interface transformer but can appear totally ungrounded from the perspective of the primary side. The grounding requirements for DR will depend on the type of system. We have discussed four-wire, multi-grounded neutral systems and three-wire, ungrounded systems. Other system types also exist such as three- and four-wire, uni-grounded systems (grounded only at the substation) and three-wire, high-impedance grounded systems. The vast majority of systems in use are of either the four-wire, multi-grounded neutral or threewire, ungrounded or high-impedance, grounded type. For these installations, Table 4-1 provides recommendations for the type of grounding to employ with DR for the most common system configurations.
Table 4-1 Grounding Recommendations for Distributed Resources (DR) Design of Primary Distribution System Design of Secondary System
4-wire, Grounded Three-wire, ungrounded system or high-impedance, grounded system 3-Wire, Ungrounded

DR Grounding Practice
DR should be ungrounded or high-impedance grounded with respect to primary and effectively grounded with respect to the secondary system DR should be ungrounded or high-impedance grounded with respect to primary and secondary system DR should be effectively grounded with respect to the primary and secondary system DR should be effectively grounded with respect to the primary system and ungrounded or highimpedance grounded with respect to the secondary system

4-Wire, Grounded Four-wire, multi-grounded neutral system 3 Wire, Ungrounded

The term effective grounding has been used in Table 4-1. IEEE standards describe effective grounding as a ratio between various sequence-impedances of the three-phase power source5. The practical effect is that a power source that is effectively grounded will not experience a voltage rise on unfaulted phases of more than 125 percent of the nominal phase-to-neutral system
5

Effective grounding means that the generator and its transformer combined as a source looking into the utility system have the following impedance characteristics: Positive sequence impedance greater than zero sequence resistance (X1> Ro) Zero sequence reactance is less than three times the positive sequence reactance (Xo <3X1)

4-27

The Impacts of Distributed Resources on Distribution Systems

operating voltage when one of its phases become faulted. The purpose of this is to limit serious overvoltages that may damage loads and equipment during faults. The concept of effective grounding and use of sequence components is complicated for persons unfamiliar with sequence component analysis and so a slightly incorrect but simplified way to explain it is shown graphically in Figure 4-13. This approach is easier to understand and provides the essence of the issue.

Figure 4-13 Example of Effective Grounding

Figure 4-13 shows a three-phase power system that is grounded at the substation. Before the conductor drops to the earth and creates a fault, the voltage vectors between phases and earth are all essentially the same in magnitude and displaced by 120o (see pre-fault vectors). The earth potential at point (N) on the diagram with respect to the substation ground point is essentially 0 volts before the fault. During the fault, the conductor of phase C touches the earth and current flows through the earth impedance back to the substation ground. Voltage drop along the earths impedance (the groundpath-return impedance) causes a shift in voltage at the earth point (N). This increases the voltage magnitude on the two, unfaulted phases at this location. The larger the return path impedance, the greater the voltage rise between phase and earth on the unfaulted phases. Proper control of the ratio of the outbound impedance path relative to the ground return path can help limit the voltage rise. The Figure 4-13 example was greatly simplified but provides an easy-to-understand explanation as to why voltage between phase and neutral/earth goes up on unfaulted phases when a single, 4-28

The Impacts of Distributed Resources on Distribution Systems

line-to-ground fault occurs. Note that on a four-wire system, there is a neutral in addition to the earth path that helps share the ground return current and thereby limit the return-path impedance. The four-wire, multi-grounded neutral system has a lot of advantages in cost, safety, and performance over three-wire systems, but it does have the drawback that equipment is phase-to-neutral connected and therefore may be more susceptible to the voltage rise that occurs during faults. Power system engineers carefully design the distribution system (if it is four-wire, multi-grounded, neutral-type system) to limit the voltage rise on the unfaulted phases by employing effective grounding in the design of the system. Effective grounding ensures that the ground return path impedance is not particularly large relative to the outbound impedance of the distribution system. This is done by using sufficiently large neutral conductors and a sufficient number of grounds as well as the correct type of transformer grounding at the substation. The system, looking out towards all points from the substation, is designed to act as an effectively grounded source, no matter where the fault is on the distribution system primary. Note that this design practice assumes that the substation or alternate feeds from other stations in emergencies are the only sources feeding into the distribution systemthe design has not accounted for DR sources feeding in from customer sites. Effective grounding means that faults to the neutral will not cause a large shift in the neutral wire potential near the fault point. This is important for four-wire designs where many transformers and other equipment are phase-to-neutral connected and damage may occur if the voltage rises too much. With effective grounding designs, the voltage on unfaulted phases with respect to earth or the neutral does not rise more than approximately 25 percent greater than the nominal system voltage level during faults. DR connecting to the system must also be designed with this effective grounding approach in mind so that it does not cause a voltage rise more than 25 percent greater than nominal on any unfaulted phases. This is easily done if the appropriate grounding and transformer type are used. Three-wire distribution systems are either ungrounded or high-impedance grounded and operate extremely differently than the four-wire system just discussed. On a three-wire, ungrounded or high-impedance, grounded system, there is no neutral return conductor and all loads get their power from phase-to-phase connected distribution transformers. The idea of effective grounding is not employed or even desired on this type of system. A fault to earth of a single phase raises the potential of the earth to that of the phase it is in contact with, so that the phase-to-earth voltage on the remaining two phases rises by a factor of 1.73 compared to the prefault conditions (see example shown in Figure 3-14). This sounds negative but it doesnt matter because loads are connected to phase-to-phase and the phase-tophase voltage does not change during the fault. In the vector diagram of Figure 4-14, note that before the fault, the phase-to-earth voltage on any phase, is approximately 7967 volts and the phase-to-phase voltage is 13,800 volts. During the fault, the voltage on phases A and B relative to earth becomes 13,800 volts because the earth potential shifts to that of phase C. The phase-to-phase voltages are unaffected. 4-29

The Impacts of Distributed Resources on Distribution Systems

Because transformers for this system type of system are phase-to-phase connected, the customer loads will not see the rise and will not be in danger. On three-wire, ungrounded systems, utility equipment that is phase-to-earth connected (such as lightning arresters) is already specified to handle any neutral shifts that occur and so is rated to withstand full, phase-to-phase voltage. For DR connected on such three-wire systems, it is desirable to limit any fault currents to earth so DR should not be effectively grounded and in fact should have no grounding or a high impedance ground with respect to the power system primary. Some form of ground fault detection to trip the unit is also desired. This case is just the opposite of what is desired for a four-wire system where we want good grounding of the DR! This is an excellent example of why it is important to know the type of distribution system to which the unit is being connected.

Figure 4-14 Example of a Three-Wire, 13.8-kV Ungrounded System During a Fault to Earth

Ground Fault Overvoltage on a Four-Wire System If ungrounded generator sources or generator sources that dont meet effective grounding characteristics feed power into four-wire, multi-grounded-neutral distribution systems, then there is a danger that an overvoltage will occur between phase and neutral on any unfaulted phases during a line-to-ground fault. This occurs because of neutral shift as discussed in the previous section. In Figure 4-15, a case is shown where a four-wire system experiences a faultthe phase-C conductor has fallen on the neutral of the power system. In this case, the substation circuit breaker quickly detects the fault and trips open. Once the breaker trips open, the DR that is connected with an ungrounded, delta-transformer winding will no longer have the substation grounding as a limiter to hold the neutral shift to a minimum amount. Instead, the neutral shifts 4-30

The Impacts of Distributed Resources on Distribution Systems

to the same potential as the phase-C voltage. We could say that when the substation breaker trips open, the system changes from a four-wire, multi-grounded neutral type to a three-wire, ungrounded design fed by just the DR. This is just like the case of Figure 4-14, but now transformers are serving various customer loads that are phase-to-neutral connected. These customer loads will suddenly experience phase-tophase voltage that is 173 percent higher than the pre-fault phase to neutral voltage. Damage will occur and a potential safety hazard also exists. Utility equipment such a lightning arresters will also likely be damaged. For such a large voltage rise, it takes only a few cycles of overvoltage to potentially damage customer loads and utility system equipment. Some utilities have taken the approach of Figure 4-16that is to allow ungrounded generators to connect, but to do so they must have a fast-acting, fault-detection scheme like that shown that trips the unit offline as soon as a ground fault is detected. The problem with this scheme is that it is extremely sensitive to routine feeder disturbances and will nuisance trip periodically. To make it less sensitive to disturbances requires that some time delay be added that could allow overvoltages to persist for many cycles and this could result in equipment damage. The only way to really avoid serious overvoltages is to design the DR system as an effectively grounded source.

Figure 4-15 Example of a Four-Wire System Experiencing a Fault The Overvoltage on the Distribution Circuit Is Fed by an Ungrounded Generator (With Respect to Utility Primary) During a Ground Fault. Note That the Generator Ground Connection on the Wye Side of the Transformer Does Not Make the Delta on the High-Voltage Primary Side Appear to Be Grounded.

4-31

The Impacts of Distributed Resources on Distribution Systems

Figure 4-16 Use of a Ground-Fault-Overvoltage Detection Scheme to Trip an Ungrounded Generator and Limit the Duration of Overvoltages

Small Generators and Ground Fault Overvoltages This discussion of grounding so far has not dealt with the size of the generator in terms of its ability to cause a serious ground fault overvoltage. Certainly, a large, poorly grounded generator such as 5 MW or larger poses a large risk to the system compared to a small generator of 100 kW or less. Whereas a large, 5-MW, ungrounded generator could create an overvoltage over a wide areaperhaps the whole feeder, a smaller, 100-kW generator could not create a sustained, serious overvoltage on a large portion of the system no matter how poorly it was grounded because there is too much load for it to drive in that scenario. On the other hand, because the potential operation of switches and interrupting devices could isolate a smaller generator with only a small portion of the feeder load that it could sustain, then even a small unit could still pose a threat to a small part of the system. Also, the aggregate impact of several small generators that are poorly grounded can pose the same threat as a single, large generator. These factors mean that regardless of size, generators should still be grounded properly with respect to the power system. Also note that the secondary system can experience some of the same type of overvoltage problems as the primary, if the grounding of the DR is not properso proper grounding can help prevent overvoltages and damage right at the DR facility!

4-32

The Impacts of Distributed Resources on Distribution Systems

Transformers and Grounding As was stated previously, the type of transformer employed has an impact on the grounding perceived by the utility primary system. For the generator to appear as a grounded source to the utility primary distribution system, the transformer must be able to pass a ground path from the low-voltage to the high-voltage side. (This is called a zero-sequence path in power engineering terminology.) Figure 4-17 shows four, commonly employed arrangements for transformer windings. Only the top two arrangements shown can provide a grounding path to the primary. Furthermore, for the transformer with grounded wye to grounded wye, the generator neutral must be grounded to make the source appear as grounded. The top two arrangements are preferred for four-wire, multi-grounded neutral systems. The bottom two arrangements shown act as ungrounded sources and are best used on three-wire, ungrounded distribution systems. An important point is that a DR site can be configured to act as a well-grounded source on the low-voltage side of the transformer, but the system may still appear to the utility primary to be ungrounded on the high side. Delta connection on the high side and grounded-wye connection on the low side can achieve this effect (see the third configuration shown in Figure 4-17).

Figure 4-17 Four Commonly Used Transformer Arrangements and the Way They Appear From a Grounding Perspective to the Primary Utility Distribution System

4-33

The Impacts of Distributed Resources on Distribution Systems

Use of Existing, Three-Phase Transformer In cases where an existing three-phase customer plans to install a DR, a desire often exists to connect the DR to the facilitys low voltage bus and use the existing distribution transformer as the interface to the primary distribution system. This saves the cost of a new transformer and much other equipment. In most situations this is fine, however, at some customers the existing transformer does not provide the necessary grounding path and the source wont appear as effectively grounded to the distribution system primary. A good example of this situation is a distribution transformer (with delta connection on the high side and grounded-wye connection on the low side) serving an existing commercial customer. In this case, the transformer provides excellent grounding for power flowing from the primary to the secondary, but if we reverse the situation and try to have power flow from secondary back to the primary, then it provides no grounding at all! This is an excellent example of the existing distribution system equipment not being designed for the flow of power from loads out to the system. Protection to prevent export of DR power to the primary system can prevent a steady state flow of power, but during transient fault conditions, the grounding problem still exists and is still an issueDR must be grounded properly whether they are designed for export or not! Transformers for Single-Phase Distributed Resources Single-phase DR, such as residential, photovoltaic systems or fuel cell systems, are normally connected to existing, single-phase distribution transformers that would have the proper grounding connections, and so it is highly unlikely, compared to a three-phase installation, that the wrong type of transformer would be present. The only possible exceptions to this could be cases where the single-phase distributed transformer was phase-to-phase connected on a four-wire, multi-grounded, neutral-type system or where a single-phase DR is used on a three-phase transformer that is not configured to provide a grounding path to the primary. Situations where these two possibilities become an issue are expected to be rare. Nonetheless, DR system integrators should be aware of the possibility. A typical, small inverter installation with proper grounding is shown in Figure 4-18.

4-34

The Impacts of Distributed Resources on Distribution Systems

Figure 4-18 A Typical, Small (<10 kW) Inverter Installation With Proper Grounding Serving a Residential Load

Circulating Currents in Transformer Windings Transformers with grounded-wye connection on the high side and a delta connection on the low side provide a ground path for a DR source connected to the utility primary, thereby providing excellent protection against overvoltages during line-to-ground faults. However, a drawback of this arrangement is that zero-sequence currents circulate in the delta winding (see Figure 4-19). To understand this, first consider that the currents and voltages of a three-phase power system can be broken down into three sets of vector components known as positive, negative, and zero-sequence components. This is a standard analytical approach that power engineers use to model power systems. The zero-sequence components are basically the ground-return-path components of current and voltage. Because of the phase relationship of the zero-sequence currents, they have a tendency to circulate around the delta winding of a grounded-wye to delta transformer in any steady state, zero-sequence component of voltage exists on the wye side of the transformer. If unchecked, these can overheat the transformer and may require derating of the transformer. Circulating

4-35

The Impacts of Distributed Resources on Distribution Systems

currents occur because the zero-sequence voltage induced on each leg of the delta winding is all in phase. Because the delta is a closed loop, these voltages will create circulating currents in the loop. Typically, a three-phase distribution transformer with grounded-wye and delta connections may have 3 percent impedance to zero sequence currents, and it is not uncommon that a 2 percent component of zero-sequence voltage on the primary exists. This would create a circulating current equivalent to 2/3 or 66 percent of the transformers current rating! This would leave only 33 percent of the rating for power flow. To solve this problem, a grounding impedance can be inserted in the neutral connection of the transformer to limit zero sequence circulating currentsthis is the common practice that works successfully for many DR sites. The grounding impedance also helps limits certain fault modes and reduces the impact of the DR site on utility protectionso it has many benefits.

Figure 4-19 Circulating Current in Delta Winding Because of Zero-Sequence Voltage

Impact of Grounded, Distributed Resource on Feeder Fault Sensing


While it is appropriate and desirable to have grounded DR to avoid overvoltages, some grounding arrangements can contribute to and interfere with the flow patterns of ground return currents that should normally flow back to the substation. A transformers with grounded-wye connection on the high side and delta connection on the low side, is an excellent example of a transformer arrangement that can cause this type of interference. The impact can be to make it difficult for relays at the substation to detect ground faults out on the distribution system. It also can interfere with fuse-breaker coordination and other protective

4-36

The Impacts of Distributed Resources on Distribution Systems

devices. Utility distribution engineers are careful as to how these transformers are connected to the system to avoid these problems. To understand how a transformer can interfere with ground fault return currents involves a discussion of the sequence current components. As was discussed in the last section, any significant zero sequence voltage imposed on the transformer grounded-wye side will result in circulating current flow in the delta winding. During a nearby line-to-ground fault a zero-sequence voltage is developed. This results in a circulating current in the transformer delta windingthis current is in a way diverted from the substation path and is never seen by the substation protection relays. This has the effect of desensitizing the utility relaying to ground faults relative to the actual currents occurring out at the fault location. It is important to recognize that this effect occurs as a result of the transformer with groundedwye and delta connections and is essentially independent of the DR connected to that transformer. A partial solution to the problem is the neutral grounding impedance. Use of a neutral grounding impedance in the ground path can limit the zero-sequence contributions (see Figure 4-20) and reduce this impact. The ground impedance needs to be sized so that it is large enough to reduce contributions but still small enough that effective grounding of the source can be maintained. This is only a partial solution because currents can only be limited so much, before the machine loses it effective grounding.

Figure 4-20 Neutral Grounding Impedance Used to Limit Circulating Current and Ground Fault Desensitization and to Reduce Ground Fault Contributions From the DR Itself

Do all DR need neutral grounding impedance? A single, small (e.g., 30-kW) DR will not significantly impact feeder protection because its zero-sequence impedance will be extremely large compared to the distribution system zero sequence impedance. In such an example, most of 4-37

The Impacts of Distributed Resources on Distribution Systems

the fault current (>99 percent) would still flow back to the substation and be measured by the relays. However, if the zero-sequence impedance of the DR source (or aggregate grounding sources) begins to approach the substation, zero-sequence impedance, by simple current divider laws, the ground fault current at the fault location will begin to be significantly different than the current measured by relays at the substation. As a rule of thumb, when the DR grounding source, zero-sequence impedance becomes less than 10-20 times the substation zero-sequence impedance, then this begins to cause significant impacts on protection (meaning a 5-10 percent difference in the measured current at the substation versus actual fault current at the fault). Larger DR installations are most likely to cause problems. Of course, a large aggregate number of small DR sources could also lead to this problem. Use of a grounding-impedance in the neutral connection of the transformer can reduce this impact to more manageable levels but an overall upper limit still exists on the aggregate capacity of DR that could be interfaced without causing difficulties. To solve any difficulties, it could require changes in relay settings and additional protective devices and modified protection schemes (by both utility system and DR). The neutral grounding reactance may be on the high-voltage or low-voltage secondary side of the power system, depending on the types of transformers used. A common transformer arrangement found at many DR sites uses two transformers to get from the generator to the utility primary; an interface transformer between the DR and the facility bus and then the existing utility distribution transformer (see Figure 4-21.) For that arrangement, the grounding reactor is installed on the 480-volt side of the system as shown.

Figure 4-21 A Common Transformer Arrangement Found at Many Commercial or Industrial Distributed Resource Sites

4-38

The Impacts of Distributed Resources on Distribution Systems

Another grounded DR installation is the installation with grounded-wye and grounded-wye connections (see Figure 4-22). This too can interfere with ground faults and a grounding reactor can help. In this case, the neutral grounding impedance will be located at the generator itself. Most rotating DR can be damaged by the severe, ground-fault current levels that occur during faults, so use of a grounding reactor or resistor on the generator is often needed anyway to prevent damage to the generator.

Figure 4-22 Grounding Impedance at the Neutral Terminal for the Rotating Generator

Single-Phasing of Transformer Banks with Grounded-Wye and Delta Connections With the transformer arrangement using grounded-wye and delta connections, if one or two phases of the high-voltage-side become disconnected, as shown in Figure 4-23, then the delta winding can magnetically couple power onto the de-energized phases. This could happen when a fuse operates upstream, a wire is broken or a switch is opened on only one or two phases. Under this condition, the transformer will attempt to supply the load on the de-energized phases by using power from the energized phase. This poses a threat because the voltage levels and phase relationships created by this mode of operation are improper and could damage loads downstream of the site. Furthermore, the load seen by the transformer in this mode of operation would be quite large compared to the rating of the transformer and could cause thermal damage. It is important to recognize that the energy that is causing the overload is not coming from the DR, rather it is coming from the one remaining energized phase and being fed back around into the other phases by magnetic coupling in the DR transformer. So tripping the DR on the low-voltage side of the transformer does nothing to stop the problem. The transformer itself needs to be tripped off line on the high side to stop the overload. The fuses often used on transformers would not necessarily melt for this type of overload and so other protection may be required to prevent thermal meltdown. This is why many larger DR employ 4-39

The Impacts of Distributed Resources on Distribution Systems

three-pole switching devices to isolate the high-voltage side of the transformer when the DR trips offline. This problem can occur at any site that has a grounded-wye connection on the high side and a delta connection on the low side and it is not limited to just DR installations. In fact, most of the documented cases are at regular customer sites where this type of transformer has been used and there is not any DR installed. In the case of DR, though, it becomes an important issue because one of the most commonly recommended transformer configurations for DR grounding purposes is the grounded-wye to delta. So if DR becomes extremely popular, there could be many more of these units out on the system. A solution to the problem is that a three-pole circuit breaker or other suitable switching device should be employed to fully de-energize all phases of the transformer if a loss or severe unbalance of voltage is detected on any phase. It should be noted that this type of switchgear is a requirement of the National Electrical Code (see Article 705-21) for such generator installations anyway, and so, if it is not present, it is a violation of that code.

Figure 4-23 A Delta Winding on the Low Side Can Result (1) in the Distributed Resource Unit BackFeed of Two De-Energized Phases, (2) in the Transformer Overloading, and (3) in Improper/Unsafe Voltage Conditions and Phase Relationships on the Feeder

Transformer and Grounding Issues that Impact the Distributed Resource Itself The discussion of transformer interface and DR grounding has focused on the utility system impact issues. However, it is important to recognize that many of these issues are critical to the protection of the DR itself. Selection of an inappropriate transformer or grounding scheme for a DR can lead to overheating of the generator and damage to the generator during some types of 4-40

The Impacts of Distributed Resources on Distribution Systems

fault conditions. There also can be damaging overvoltages to the loads at the DR site if it is incorrectly grounded. Careful attention to the proper design details of grounding, protection and transformer interface not only help the utility company avoid problems but also avoid damage and dangerous conditions at the DR site. Having stated the above, this is not to say that there arent some design needs of the utility system and DR that are in opposition. One example is the issue of grounding of the DR source with respect to the utility system primary. Most DR designers, if they were given a choice, would rather have an ungrounded or high-impedance grounded interface with respect to the utility primary distribution system. This can help reduce the impact of ground faults on the generator and eliminates the circulating currents that can occur in some types of transformer configurations. Fortunately, there are design approaches that can generally satisfy the grounding needs of the utility and also meet the DR needs. While these may not be the ideal arrangement for either party alone, they serve as a technical compromise that will satisfy safety and operating requirements. General recommendations are discussed in the next section. Overall Conclusions on Transformers and Grounding The issues discussed in the preceding sections show that there are many factors that must be considered in the selection of appropriate transformers and grounding for DR. While there is still not 100 percent technical agreement in the industry as to the best transformer arrangements, a general technical consensus exists that effectively grounded configurations should be used to interface DR with four-wire, multi-grounded neutral systems. Both the transformer with grounded-wye connection on the high side and delta connection on the low side and the transformer with grounded-wye and grounded-wye connections can be suitable to provide an effectively grounded interface (note that in the case of the grounded-wye to grounded-wye connections, the generator neutral must not be grounded for this to be true). However, these configurations should usually be properly equipped with neutral grounding impedances to avoid the issues that have been discussed in earlier sections. For three-wire, ungrounded distribution systems, the DR should not provide a ground path for current flow to ground or at a minimum should have a high impedance path that can be quickly tripped off line. An ungrounded transformer interface (such as a delta to delta) or a delta connection on the high side to a grounded-wye connection on the low side can satisfy this objective. If a grounded-wye transformer is to be used on the primary side of the three-wire system, it should only be used if it is grounded through an extremely high impedance. In all cases, the DR should be equipped with ground fault detection that trips the unit off line upon occurrence of a ground fault. For ungrounded installations, this could be the broken delta PT method shown earlier in Figure 4-15. For grounded systems, standard current sensing methods could be used. Table 4-2 summarizes the transformer and grounding recommendations. Because of variations in the distribution system design and protection practices amongst the many different utilities in the 4-41

The Impacts of Distributed Resources on Distribution Systems

country, be advised that Table 4-2 is only a guide for most utilities and that some utilities may need different arrangements to ensure safety and reliability of their system. Also, in some instances, several transformers may likely exist between the DR and utility systemall transformers cascaded in series should provide the appropriate grounding connections with respect to the part of the system to which they are connected, as well as the final interface to the primary. As a public policy matter, it is possible to standardize transformer selection for most smaller DR applied at residential sites or small commercial sites on feeders, where total penetration of DR is low. However, at larger industrial and commercial sites, a certain amount of custom engineering will be needed to get around problems that can arise.
Table 4-2 Recommended Transformer Configurations for Three-Phase Distributed Resource Installations Winding Configuration Employed for Interface Between the Generator and Utility System Utility Primary Side Generator Side (Low voltage side) Is it a Grounded Generation Source with Respect to Utility System Primary?

Type of Distribution System to Which this Transformer Should Normally Be Applied

Yes

Suitable for four-wire, multi-grounded neutral systems. It may need a neutral grounding impedance and three-pole switchgear on the primary of the transformer
Suitable for most four-wire, multigrounded neutral systems but the generator must be grounded. May expose the generator to severe fault forces neutral grounding impedance may be needed.
Not usually recommended for four-wire, multi-grounded systems. A good choice for three-wire, ungrounded distribution systems. Not usually recommended for four-wire, multi-grounded neutral systems. A possible choice for three-wire, ungrounded distribution systems.

Only if the generator neutral is grounded

No

No

4-42

The Impacts of Distributed Resources on Distribution Systems

Service Reliability
Utilities generally measure the reliability of power delivered to their distribution system customers by tracking service interruptions that impact their customers. Two types of service interruptions on the distribution system that utilities are particularly interested in tracking are sustained interruptions and momentary interruptions. Generally, sustained interruptions are defined as those lasting 5 minutes or longer. Momentary interruptions are those lasting less than 5 minutes (often these may be just a few seconds). Depending on the type of distribution system and the way it is protected, customers will experience various combinations of these two types of service interruptions. Minimizing the number of momentary and sustained interruptions experienced by various customers is extremnely dependent upon the proper functioning of the grid protection equipment during fault conditions. DR can negatively affect service reliability because, as has been discussed in earlier sections, it can confuse the operation of grid protection equipment including fuses, reclosers, circuit breakers and other devices. If the equipment does not function properly, utility customers can experience longer and more frequent outages.

Islanding of Distributed Resources on Radial Systems


Islanding is the situation in which a DR installation and a portion of the utility system have become isolated and the DR continues to operate and serve loads on the circuit. Islanding can be intentional or it can be unintentional depending on the objectives and controls employed. Unintentional islanding into any part of the utility system can compromise safety and can harm customer loads and distribution system equipment. Some key dangers of islanding include: Island will drift out-of-phase with respect to the utility system and then equipment may be damaged during utility company reclosing. Incidents of energized, downed conductors may increasecausing more public danger. Service restoration by utility crews may be delayed and made more dangerous. Islands may not maintain proper voltage or frequency for connected loads this can threaten customer loads and utility equipment. Ferro-resonance or incompatible grounding practices of some DR that dont show up during normal system operation may lead to damaging voltages during the island existence.

A variety of island scenarios may occur depending on the size of the DR, configuration of the power system and load that exists on the system. A large DR, such as a 5-MW unit, has enough power output capability that it could potentially island an entire distribution feeder, whereas, a small DR, such as a 10-kW PV system, could island only a group of houses at most. One requirement for islanding is that the DR device must be capable of self-excitation, which is true for synchronous generators and force-commutated inverters. Generators that normally are not self excited, such as induction generators or line-commutated inverters are generally not able to island. However, they can island under some conditions. 4-43

The Impacts of Distributed Resources on Distribution Systems

Figure 4-24 shows an example of a large island that might be formed if a 3-MW synchronous generator is isolated with a major portion of the distribution feeder. Figure 4-25 shows a potential, small island on a secondary serving just a few customers.

Figure 4-24 Example of a Large Island That Might Be Created if a Large Distributed Resource (e.g., 3 MW) Is Isolated With a Major Portion of a Distribution Feeder

4-44

The Impacts of Distributed Resources on Distribution Systems

Figure 4-25 Example of a Small, Unintentional Island on the Secondary System if the Fuse Should Operate or Be Removed and the Anti-Islanding Protection in the Local Generators Does Not Work Properly

Once an island forms with DR, the utility system no longer controls frequency or voltage in that section. It is the DR that is controlling the voltage and frequency and it would not ordinarily be equipped to handle that function over a large area. Customers may be subjected to voltages and frequency that damage equipment or imperil human safety. Islanding may also damage utility equipment and cause service restoration delays because, once established, an island will typically be out of phase with the utility voltage. To reestablish normal operating conditions and rejoin the islanded section of line to the main circuit, the two must be in phase and this coordination takes time. Rejoining the island in phase is imperative, as an out-of-phase connection could damage the DR or utility equipment, which is not rated for that capability. Utility personnel who are working near equipment that fails because of out-of-phase reclosing could be seriously injured. The effects of islanding may also pose a public safety risk because there could be an increased hazard of downed conductors that can be touched by the general public and the protection system is less likely to successfully de-energize lines that have fallen onto automobiles, energized other public services, and perform other protective activities (see Figure 4-26).

4-45

The Impacts of Distributed Resources on Distribution Systems

Figure 4-26 A Distributed Resource With Improper or Ineffective Anti-Island Protection Could Pose a Threat to the Public and Utility Workers by Back-Feeding Downed Conductors

Fortunately, most of the potential cases of islanding can be prevented if DR use appropriate anti-islanding protection. The most common means of preventing islanding is what is often referred to as the passive, anti-island protection technique. This approach uses a voltage and frequency relay set to trip the DR whenever either parameter migrates outside an acceptable window. Relays would typically be set to a tight frequency range on the order of 0.5 to 1.0 Hz around the nominal frequency of 60 Hz in the United States. The range of allowed voltages are a bit wider, typically 5 to 10 percent, to allow for normal variations. Undervoltage, overvoltage, and frequency protection equipment prevents islanding because in most cases when a section of the distribution system and DR unit separate, the output of the DR unit will not match the power demand within the separated area. For synchronous or induction generators, this will result in a change in voltage and frequency that causes the relays to trip in an extremely short time frame. An IEEE standard under development (Std. P1547) is currently recommending that DR cease export of power within 10 cycles for a severe, abnormal, low-voltage conditions (less than 50 percent). These would be indicative of an open switching device upstream or nearby fault. Because the utility company circuit breaker may attempt to reclose within anywhere from 12 cycles up to more than 90 seconds, the recommended anti-island practice uses 10 cycles because no standard equipment exists that can reclose faster. 4-46

The Impacts of Distributed Resources on Distribution Systems

Figure 4-27 shows the concept in action. It should be noted that another key part of antiislanding protection is the blocking of reconnection to the utility system. This means that once the relays trip the DR off line, the control logic is programmed to prevent any attempt to reconnect into an out-of-range utility system voltage or frequency condition. Only after voltage conditions have been restored to the normal range is DR allowed to reconnect. Note that some utilities use this blocking approach on their circuit breakers in substations if there is concern that DR may be islanding on the power system.

Figure 4-27 Illustration of the Concept Used in IEEE Standard P1547

In situations where the load on the island is nearly balanced with the generation on the island, the passive technique may not be adequate because there may not be sufficient voltage or frequency change to cause the unit to trip. This is particularly true for inverter-based technologies that can rapidly respond to changes in load and have extremely stable frequency output. To overcome this problem, more robust, active, anti-islanding techniques have been developed for many inverters. These approaches employ a combination of active and passive techniques including: Overvoltage and undervoltage detection, based on the assumption of a mismatch between reactive sources and loads in island Overfrequency, underfrequency, and rate-of-change of frequency detection, based on the assumption of a mismatch between active sources and loads in island Harmonic content detection, based on higher relative harmonic levels for islanded systems containing high percentages of inverter sources Comparison of the response of the DR unit to changes in power set point (response will be faster when grid connected) 4-47

The Impacts of Distributed Resources on Distribution Systems

Active frequency shift, based on intentional frequency shift of inverters away from normal frequency using positive feedback Active voltage shift, based on positive feedback of voltage changes Active power shift, based on detection of current and/or frequency changes during intentional power modulation System faultlevel monitoring, based on measurement of the power systems source impedance close to the interconnection point

Widely used active techniques are the voltage- and frequency-shift techniques. In those cases as soon as the island forms, the DR does not remain stable, and its frequency or voltage begins to shift until eventually it shifts outside of the acceptable window of its passive voltage and frequency protection relays. To be highly effective, this sort of shift must occur rapidly once the unit becomes islanded but the unit must be stable enough to remain locked to the utility system voltage and frequency while the reference utility voltage is present. A typical requirement for this protection is that it should operate within a few seconds following the isolation of DR from the grid for a balanced load island scenario or a minor voltage or frequency excursion outside the limits. For more severe situations, such as a deep voltage sag, it should operate within 10 cycles or less. Simulations and tests of passive and active techniques with different load types revealed that some active technique detection times could be longer than the requirement if multiple inverters are connected and the load type meets certain criteria. Further study is needed to evaluate cases in which islanded systems are formed to comprise mixed forms of generation including synchronous sources. So far though, with inverters, tests confirm that frequency- and voltage-shift techniques both generally provide rapid detection when used even with worst-case loads. IEEE Standard 929 and UL Standard 1741 define anti-islanding test procedures that represent some of the worst-case scenarios that are expected to occur in practical situations. Many DR operators desire to operate their units in an islanded mode as a source of standby power during a utility interruptionsupporting only their own load (see Figure 4-28). Utility companies have no issue with this approach as long as the units are not islanded with any portion of the utility system and have a proper means of being resynchronized and connected in phase with the utility system once power is restored. DR equipment exists that has been developed to provide anti-island protection for the utility system connection, but can continue to operate as a tiny island supporting only site load for the DR customer. It also is possible to have interlocks that bypass the anti-island relays on a DR only after a utility disconnect switch has been opened. This allows the DR operator to pull the switch open, isolating the DR from the utility system, and then restart their DR to serve only the DR site load.

4-48

The Impacts of Distributed Resources on Distribution Systems

Figure 4-28 Anti-Island Protection Can Be Arranged as Shown Here to Isolate the Distributed Resource From the Utility System But Continue to Operate a Local Customer Facility

Intentional Islanding While we have discussed the dangers of islanding of DR on the utility system, it is important to recognize that islanding may improve reliability if it is done as part of a carefully planned program with the proper controls and equipment. When done in this manner, it is called intentional islanding. This is an option that many utilities are considering but it has not seen much deployment yet. It must be implemented properly and with great care to avoid causing the safety and operational concerns that we have already discussed. An example of intentional islanding is shown in Figure 4-29 where a DR has been strategically placed downstream of a recloser. During an upstream fault, it could improve service reliability for the customers served on the tail end of the circuit by islanding and continuing to support those customers (the recloser would open to provide isolation of the island). To be successful, the DR must be able to support the islanded load, maintain adequate voltage and frequency, and handle any transient-starting inrush required to restart the island (should the island be dropped by the DR). This all needs to be accomplished while ensuring that the proper protection and communication is provided to control the recloser and coordinate the operation of the DR. The recloser should be blocked from reconnecting to the utility system until the utility system is re-energized at the proper voltage and frequency and is in-phase with the island. This scheme could greatly benefit rural systems by providing an alternative source in areas that have no such option. It is not likely to be done with customer-owned DR but rather would be part of a strategic use of utility-owned or energy service company DR to improve system performance. 4-49

The Impacts of Distributed Resources on Distribution Systems

Figure 4-29 An Example of Intentional Islanding Where a DR Has Been Strategically Placed Downstream of a Recloser

Avoiding Islanding on Network Systems and Interactions with Network Protectors Secondary network distribution systems function quite differently from radial distribution systems and have correspondingly different complications with the interconnection of DR. One complication arises from the use of network protectors and the potential that an island will develop on a low-voltage network. To understand this danger, lets first discuss the operation of network protectors. The purpose of network protectors is to prevent reverse power flow from the low-voltage network into a faulted network transformer or primary feeder cable (see Chapter 3 for details on this function). These devices have sensitive, reverse-power relays that will trip on a tiny amount of reverse power flow (much less than 1 percent of the normal forward flow). If care is not exercised in the application of DR on the low-voltage network, then DR can trip such protectors and island the low-voltage network from the main utility system. This is particularly dangerous and should be prevented by limiting the amount of DR on the network and/or making appropriate adjustments to the network and DR protection equipment so that this will not occur. Figures 4-30 to 4-32 depict the connection of a secondary, spot network to several primary feeders and show the operation of the protectors under normal conditions, faulted conditions, and with a large DR connected. In Figure 4-30, the network is fed by three feeders that , under normal operating conditions, feed a nominally equal level of power into the network. The 4-50

The Impacts of Distributed Resources on Distribution Systems

resulting redundancy in power supply lends networked systems their reliability, but also makes them more complicated to protect. When a fault occurs on one of these feeders, as shown in Figure 4-31, power feeds into the faulted feeder from both the substation side and the low-voltage network side (by way of the network transformer). To de-energize the fault requires that the substation breaker and the network protector both trip. Power feeding from the low-voltage side will be in the reverse direction and the network protector is set to trip on an extremely low value of reverse power (less than 1 percent of the normal forward rating). Because of the redundancy of the network system design, with the faulted feeder removed from the system, the remaining two feeders have adequate capacity to supply the load on the network.

Figure 4-30 Low-Voltage, Spot Network Under Normal Operating Conditions

4-51

The Impacts of Distributed Resources on Distribution Systems

Figure 4-31 Low-Voltage Spot Network With One of the Primary Feeder Cables Faulted

As is shown in Figure 4-32, placement of a distributed resource with a capacity large enough to exceed the network load could cause enough backfeed to the system such that all of the network protectors connecting the feeders to the network would open. At this point, the network would be energized only by the distributed resource and the generator-fed network system would be separated from the feeder system. Several problems exist with this scenario, ranging from poor power quality to nuisance tripping of protectors to equipment failure and major hazards. One of the most critical hazards is that the network island will not stay at the same exact frequency as the utility systemthe difference in frequency between the utility system and the island is referred to as a slip frequency. While this slippage is occurring the utility system and network-island periodically drift into and fully out of phase. During a moment when they are in phase, it is possible for the network relay to be fooled into believing that conditions are appropriate to reconnect. However, by the time the reconnect command is completed, the network-island and utility system will again have drifted out of phaseso the network protector will actually close into an out-of-phase system. This will usually damage the network protector as well as the DR equipment and loads on the island.

4-52

The Impacts of Distributed Resources on Distribution Systems

Figure 4-32 Example of Distributed Resource With a Capacity Large Enough to Exceed the Network Load

As an example of this problem, suppose that a large urban hotel with a spot network has determined that a co-generating distributed resource would be cost effective. They proceed to interconnect a cogeneration unit with a capacity slightly larger than their minimum load to the spot network. On a lightly loaded day, the hotels load drops below the output of the distributed resource, and power begins to back-feed through all three of the network feeders. The network protectors trip, disconnecting the hotels spot network from the spot network bus, and causing the distributed resource to island with the load. After these systems become separated, they will drift out of phase and become unsynchronized. Eventually, the phases of the network bus and the DR-island will come close enough together to initiate the reclosing of the network protectors. However, by the time the protectors are closed, the island may have swung out of phase again because it takes approximately second to fully initiate the command. The relays that control the automatic reclosing of network protectors were never intended to act as synchronizing relays for interconnecting two, non-synchronized systems. Also, the network protectors themselves are not tested or rated for this type of interruption duty and it is still being evaluated whether this type of duty could pose a threat to most protectors. Some extremely serious incidents have existed where network protectors have burned up when apparently trying to interrupt the electric power system from a DR island. Technical solutions exist to this problem. Most obvious is that the DR power output could be limited to a capacity smaller than the buildings minimum load. This is not as easy as it may 4-53

The Impacts of Distributed Resources on Distribution Systems

seem. First of all, it is extremely difficult to decide the minimum load level that should be chosen. Is it the minimum load that occurs on a daily basis, the minimum that occurs weekly, or the annual minimum load? What is the annual minimum loadis it even known? What if a fault or overload in the building trips several circuit breakersshutting off most of the load in the building and creating a lower than expected minimum load? Clearly, to be safe, one would need to size the generator to consider the likelihood of all these scenarios and have operating procedures that would shut down the DR unit before tripping any significant loads in the building. Even if the generator could always be operated at less than minimum load under steady state conditions, this does not prevent transient reverse power momentarily flowing into the protectors if a voltage sag should occur because of a fault on the utility system. For example, if a voltage sag occurs on the utility system because of an upstream fault at the substation bus, the DR could momentarily cause reverse flow into all of the network protectors serving the spot network and thereby trip them. This, of course, would island the network. One way to be reasonably confident that the islanding problem can be avoided is to conservatively limit the size of the DR to a small value compared to the minimum load on the networkperhaps only 10 percent of the worst-case, minimum load. Even better, if the DR is limited to the size of the reverse power trip threshold of the most-sensitive network protector, then this will solve the problem altogether. Of course, with either of these approaches, the generator would end up being extremely small compared to the building loadin many cases, too small to meet the objectives of the building owner that installed it in the first place. In cases where a larger generator is to be applied, then modifications to the network protection are required. These changes include using time-delayed relaying instead of the normal, instantaneous, reverse power relaying on the network protectors coupled with an instantaneous, reverse-power relay that trips the DR when reverse power is detected. Time delays are necessary because the normal instantaneous trip on the network protector is so fast it will trip because of reverse current before most DR could be tripped off line. With a time delay on the network protectors, the DR can be set to trip off line before any network protector trips. While making these relaying changes may be one possible solution, they cost money and, more importantly, could cause other problems for the utility system. For example, the long time delays on tripping of faults would increase the amount of damage caused by faults and also degrade power quality. Also quite a bit of controversy still exists regarding the ability of network protectors to serve as an interrupting device between a DR island and the utility system. So, the IEEE P1547, Draft 7, standard does not currently recommend that they be used for such duty, unless rated for it. New technology is being considered to help resolve this issue. One proposed future approach is to use special electronic controls and communication paths to allow any network island that does form to stay in phase with the utility system. This would eliminate the danger of islands and allow the network protectors to function as they are now without any changes. This technology is not available yet but is under consideration by EPRI for future development. Manufacturers of 4-54

The Impacts of Distributed Resources on Distribution Systems

network protectors are also developing new products that can address the needs of DR on the system.

Voltage and Frequency Control


Maintaining adequate operating voltage at all customer delivery points is critical to proper system operation. As has been discussed earlier, the range of acceptable customer service voltage is normally classified as 5 percent on the nominal level, with +6% or -8% acceptable for occasional and short-term events. These levels are described in ANSI standard C84.1a national standard that all utilities must consider. It should be noted, however, that a few utilities have even tighter standards that supersede the ANSI C84.1 limitsstate regulatory commissions impose these. For example, some utilities are required to maintain voltage within a +3% to 5% band. Frequency is also maintained within extremely tight tolerances to ensure proper operation of connected equipment. On a large, interconnected power system, small deviations in the 60-Hz system frequency of as little as tens of millihertz, represent large imbalances between load and generation, so the frequency does not normally change that much. In reality, this tight frequency control is more an artifact of the large size of the national electric grid and physical needs in controlling it, as opposed to concerns over damage to connected loads, which can tolerate much larger variations. In considering the impact of DR on the power system voltage and frequency, it should be recognized that DR has a greater impact on the system voltage than it does the frequency (at current, low-penetration levels for DR). This is because it can locally change the voltage where it is applied without having to change the voltage across the entire power system. Whereas, to change the system frequency requires a system-wide impact and so the capacity of the DR needs to be significant relative to the total system capacity. The largest individual DR units (10-50 MW) are still less than 0.01 percent of eastern or western U.S. area generation, and so any single DR does not significantly impact frequency. If the output of all DR units currently connected on the system were coordinated to raise or lower the system frequency, then the impact could be noticeable, but even here it would still not be that great because aggregate DR penetration on these systems is under 5 percent. In the future, if DR penetration were to increase to 10 percent or beyond, then frequency impacts could become an issue and utility companies will need to carefully model the impacts of DR generation on the system frequency. Some of the safeguards currently employed to protect the distribution system (anti-island controls) may actually destabilize system frequency once the DR reaches a significant penetration level. While, this is not an issue now except for smaller power systems, in 10 years it could become an issue. On smaller bulk systems, such as geographically islanded power systems and other isolated locations, DR aggregate capacity with respect to system generation has in some cases already become large enough to significantly impact frequency, and problems have occurred on such systems. Fluctuating outputs from wind generation plants and other sources can push the frequency outside of appropriate limits unless the proper feedback control from generation is 4-55

The Impacts of Distributed Resources on Distribution Systems

used to stabilize the frequency. Studies are performed to assess the worst-case conditions of frequency change and implement the appropriate DR limits and/or system upgrades to prevent problems. The condition where DR has the greatest influence on frequency is where it is operating as an island in itself. It then has complete control over the system frequency. If DR is to operate as an island, then it must be able to maintain relatively stable frequency. The frequency conditions needed for proper operation of loads would vary depending on the type of loads and applications. However, it does not need to be as tightly controlled as the bulk power system and frequency within 2 percent of nominal frequency should ordinarily be adequate for islanded DR operation. Many loads are capable of operating with even broader frequency excursions. Direct manipulation of the synchronous generator inputsmechanical torque and field current applied to the rotoris the primary method of controlling the frequency and voltage of rotating generators and would be the practice for most islands. DR islands are more prone to frequency excursions during loads steps and rejections than bulk power systems because of the limited inertia of small generators and typical time lags associated with increasing or decreasing generator output to match demand. For DR connected to distribution systems at current and expected, near-term, penetration levels, voltage impacts are the main concern as opposed to frequency impacts. In general, utilities desire for the DR to have the minimum impact possible on system voltage or for it to support the voltage. The most appropriate operating practice for DR to avoid problems is to operate in a voltage-following mode. This means that the unit injects real power at the appropriate power factor (near unity) and makes no attempt to directly regulate voltage on the utility system (meaning it does not try to force the voltage to a fixed set point or range by adjusting its reactive output). This is the method recommended in most DR interconnection guidelines and the under-development standard IEEE P1457. The term voltage-following has been misinterpreted by many to mean that a DR operating in such a mode will not cause any impact on the power system voltage. However, this is not the case and it is important to recognize that a DR operating in voltage-following mode if current is coming out of its terminals will impact the system voltage. The voltage change depends on the amount of current injected, the impedance of the power system at the point of injection, the phase angle of the current with respect to the utility system voltage and any interactions with utility voltage regulation equipment. In some cases, the voltage may increase, and, in others, it may decrease. When real power is injected at the tail end of a voltage regulation zone6, the voltage will increase. Figure 4-33 is a simplified view of this hypothetical case showing the voltage profile on a feeder with several different levels of power (none, 1 MW and 2 MW). In this example, 1 MW of injected power provides some support of the feeder voltage helping to raise the voltage near
6

A voltage regulation zone is the region between the output terminals of the nearest upstream step regulator or LTC transformer device and the end of feeder or next downstream step regulator device.

4-56

The Impacts of Distributed Resources on Distribution Systems

the end of the feeder. However, if too much power is injected (see case with 2 MW), the voltage could become high at the tail end of the feeder as shown. DR also has the potential to cause low voltage on the feeder if placed near the output side of a voltage regulator with line drop compensation (see Figure 4-34). In such situations, the regulator is shielded from measuring the true load and does not raise the voltage adequately. This impact can be avoided or limited by adjusting regulator settings, but in cases where the DR is large, setting changes may not be enough and new regulation equipment may be needed or the upstream regulator may need to be moved.

Figure 4-33 Voltage Profile on a Distribution Circuit With and Without Distributed Resources

4-57

The Impacts of Distributed Resources on Distribution Systems

Figure 4-34 Voltage Profile on a Distribution Feeder with Distributed Resource Added near a Voltage Regulator Station

As DR units are added to distribution systems and aggregate capacity builds up to a significant portion of total feeder load, coordinating DR with distribution system regulators becomes extremely important. Several possible modes of interaction exist depending on the type of DR and its control configuration. In addition, the presence of DR will directly affect voltage profiles along a feeder by changing the direction and magnitude of real and reactive power flows. The directional characteristics of voltage regulation circuitry must also be considered. For cases where reverse power flows through a regulator with controls not designed to handle it, the regulator can literally run away to extreme setting positions that cause severe voltage excursions on the feeder. If DR is expected to cause reverse power flows at a regulator, then that regulator should either be capable of handling this situation or its settings should be adjusted to minimize negative impacts. In general, any time the DR capacity exceeds approximately 10 percent of the feeder load, the regulation impacts need to be studied either through load flows or other methods to ensure that voltages are still within limits. While our discussion so far has focused on the impact of DR on the primary voltage, the impact of DR output on the power system secondary also needs to be considered. The secondary capacity is much less and it is therefore more sensitive to injected power. It is possible for even a small, 10-kW generator to cause an abnormally high voltage on a secondary. Because the 4-58

The Impacts of Distributed Resources on Distribution Systems

secondary may have multiple customers, this means that a DR at one customer could impact several others. Figure 4-35 is an example of this situation whereby a secondary is exporting power back to the utility primary and this causes a voltage rise at the transformer and secondary conductors that leads to high voltage (voltage greater than ANSI C84.1, Range B, limit of 127 volts). Such conditions could occur if the site was near the front end of a regulator zone (so it had higher feeder voltage to start with), the DR output was near its maximum at the time of higher voltage on the feeder, and the loads were minimal at the customers at the time of maximum DR output. While this case is expected to be rare, the present, voltage-operating window allowed in IEEE 929-2000 and the most recent drafts of the new IEEE DR standard (IEEE P1547 D7) create an opportunity for this to occur because the DR can continue to operate until 132 volts. This problem can be avoided by not allowing DR to operate after the voltage exceeds 127 volts for an extended period of several minutes.

Figure 4-35 Injected Power From a Distributed Resource at One Customer May Influence Others on the Same Secondary

So far, we have discussed operating DR in a voltage-following mode, but, in some cases, it is technically possible to operate DR in a voltage-regulation mode by varying its reactive output to hold the voltage within a desired range. The use of DR for such active voltage regulation purposes is normally avoided and not desired by the utility company, but exceptions exist. One reason why this might be done is that the capacity of the DR is large and it has a severe impact on feeder voltage if operated in the standard voltage-following mode.

4-59

The Impacts of Distributed Resources on Distribution Systems

By regulating reactive output, the DR could help control voltage to minimize any serious voltage excursions and avoid a major retrofit of voltage regulation equipment on the feeder. Some DR power converters can operate in a partial, reactive-power control mode to help minimize the impacts of fluctuating power and are marketed with this capabilityso, equipment is already on the market offering this capability. It is also possible to regulate voltage at essentially just a DR site bus in certain configurations if sufficient upstream impedance is available such that a small amount of reactive energy can be used for that purpose. These are methods used in specific cases where appropriate. In the majority of DR installations, active regulation support of the utility system by DR is normally avoided for a number of reasons. First, on an individual basis, most DR are too small to regulate voltage on the feeder or could not spare the reactive capacity needed for this. They need to be large or there needs to be a large, coordinated aggregation of DR available for this. Secondly, the voltage control scheme must be carefully planned and in full coordination with the utility company voltage regulation equipment. If this is not done, it is possible an attempt to regulate voltage could cause more harm than good. Thirdly, this operating mode normally has negative economic consequences for any owner of DR because they would need to give up real power capacity to meet the reactive power requirements needed for voltage regulation purposes. Finally, the utility company is liable for voltage conditions on the feeder and would not normally want another entity having control of the voltage. For all of these reasons, such active control will for the time being remain the exception rather than rule. In the future, though, as technology of DR automation advances and the distribution system becomes more interactive with DR, it is quite possible that some active voltage regulation functions may be employed in most DR. The addition of DR with appropriate controls in the future provides the potential for faster and more uniform voltage regulation, leading to improved customer voltages at the point of delivery. DR could adjust its reactive output in a rapid dynamic manner to mitigate voltage flicker and other power quality conditions.

Stability
Steady state stability is the ability of the power system to transmit power under relatively stable (non-disturbance) conditions. For a simple, radial connection between a single generator and an effectively infinite system, the stability limit is related to the voltage phase angle between the machine and the power system. Simple formulas can be used to determine the maximum power that can be transmitted. However, for a multi-machine system with distributed loads, this calculation will produce unreasonably pessimistic results. The techniques are still useful to help screen cases and determine which will require a more rigorous analysis. Although steady state stability is unlikely to impose a limitation on most DR applications, large machines with high, synchronous reactances that are connected at low-voltage levels present potential problems. If such a generator cannot be located at a stronger point on the distribution systemfor example, closer to the substationthe most reasonably priced solution is to ensure that the unit is equipped with a high-gain automatic voltage regulator (AVR) and high-initial-response excitation system. This combination increases the synchronizing forces acting on the machine rotor and will thus improve transient stability response. 4-60

The Impacts of Distributed Resources on Distribution Systems

Transient stability is the ability of the generators to remain in step during sudden power system disturbances such as a fault or extremely large load rejection. Transient stability is a function of many variables: generator size and impedances, transmission strength, fault type and location, clearing time, generator excitation configuration, and settings. Transient stability limits are commonly approximated using simple, algebraic criteria. These techniques use the steady-state generator rotor angle to assess stability for different connection strengths. Note that these techniques are not recommended because more sophisticated computational tools are readily available to perform a rigorous assessment of transient stability for credible contingencies on the related distribution system. As with the steady state analysis, simplified tools will normally provide pessimistic results because they do not account for the positive effect of high-gain AVRs in increasing synchronizing torque and improving transient stability limits. The use of high-gain AVRs and high-initial-response excitation systems together with power system stabilizers can alleviate both transient and small-signal stability concerns. The addition of DR to the distribution system can affect the stability of the distribution circuit. For instance, discrepancies between scheduled load and actual load cause small load disturbances on distribution circuits. On a system without distributed generation, these disturbances merely cause momentary deviations in the power profile. However, with the introduction of DR on the system, these momentary load disturbances have an effect on not just the substation and the large transmission grid behind it, but also on the smaller DR units. Because these units do not have as much machine inertia as the generators that supply the transmission grid and its substations, they are much more susceptible to slight deviations in load and can cause frequency instabilities as a result. Instabilities can result in power flow oscillations between various DR on the distribution system and loads. These oscillations can cause voltage flicker and other problems.

Sags, Swells, and Momentary Interruptions


Sags, swells, and momentary interruptions are short-term variations in the system voltage that are usually associated with faults. The switching of large loads or the starting or stopping of motors and other large equipment can also cause sags and swells. Sags, swells and other transients have the potential to damage extremely sensitive customer equipment in some circumstances. DR can cause such transients if a large-capacity unit is turned on all at once, or a number of units are turned on simultaneously, or if a resource comes on-line before synchronizing with the grid. These events can lead to out-of-phase conditions for rotating equipment that may place undue mechanical stress on the equipment leading to damage. Sags, swells, and momentary interruptions on the distribution system can trip all DR on a given feeder off line because the DR is required to have sensitive, anti-islanding protection that intentionally trips units off line during such disturbances. If the DR is not being relied upon to support the feeder loads, then this is not an issue. However, if it is crucial for feeder support, then 4-61

The Impacts of Distributed Resources on Distribution Systems

this poses a problem. How does one maintain good anti-island protection while still helping to maintain feeder support with the DR? A solution is to have multi-level tripping where a long time delay of several seconds exists for minor voltage sags and swells and a short, tripping time delay (10 cycles or less) for more serious sags, swells, and interruptions. The IEEE 929 standard for photovoltaic interconnection has adopted this approach. Many DR installations use relays with multi-level trip functions that can handle this situation. As DR penetration increases, it will be necessary to better coordinate the tripping of DR to minimize the impact of feeder sags and swells and allow the feeder to quickly recover from such events without a long-term deficit of power capacity. Many current interconnection standards have a 5-minute-waiting period for the DR to reconnect following return to normal voltage on the feeder. In cases where the DR capacity is required to support feeder operations, this may prove to be too long and will need to be modified!

Harmonics
A harmonically distorted wave is one that does not follow a pure sinusoidal pattern as shown in Figure 4-36. Mathematically, a harmonically distorted wave shape can be broken down into its fundamental frequency (60 Hz) plus numerous other components such as 120 Hz, 180 Hz, and 240 Hz. These other frequency components are called harmonics and ideally the system performs best if they are minimized as much as possible. Harmonics are always present on utility systems to some extent. They can be caused by nonlinearities in transformers exciting impedances or loads such as fluorescent lights, AC-to-DC conversion equipment, variable-speed drives, switch-mode power equipment, arc furnaces, welders, and other equipment. Installation of DR can, in some cases, increase harmonics on a utility distribution system from acceptable to objectionable levels.

4-62

The Impacts of Distributed Resources on Distribution Systems

Figure 4-36 Comparison of Ideal Sinusoidal Wave and a Severely Distorted Wave

High levels of harmonics can lead to a variety of problems including overheating of motors and neutral conductors, resonance in capacitor banks and other system elements, nuisance tripping of protection devices and poor operation of loads that are sensitive to harmonics. In the worst-case scenarios, damage to power system equipment or loads may occur. The level of harmonics generated by a DR and its influence on the power system depends on the type of power converter employed, its design characteristics and the impedance characteristics to the power system. In general, most modern DR will not produce considerable harmonic content, if appropriately specified and installed, and so are not a great threat. Many fears regarding harmonic distortion stem from the use of early static power converters in the 1970s and 1980s. These converters were primarily based on silicon-controlled rectifiers (SCRs); the waveform created by an SCR inverter looks like a modified square wave and is extremely rich in harmonics. Today, insulated gate bipolar transistor- (IGBT-) based inverters dominate the market and have extremely low distortion because of their high-frequency switching and pulse-width modulation approach to wave generation. Most IGBT inverters can be applied in a manner that satisfies IEEE 519-1992 standards for harmonic distortion and can, in fact, produce a lower-distortion waveform than typical utility line voltage. Some rotating generation equipment can generate high 4-63

The Impacts of Distributed Resources on Distribution Systems

levels of triplen harmonics (3rd, 6th, 9th, and so forth). These can be minimized by use of a 2/3 pitch winding design. If harmonics are an issue, filters and other equipment can be applied to help mitigate them. The type of interface transformer used to connect the DR to the power system also impacts harmonics. A transformer that does not provide a generator zero-sequence path to the distribution primary will not pass triplen harmonics to the system. On the other hand, if the generator is rich in harmonics, use of such a transformer could locally make harmonics at the customer site worse. Harmonics need to be evaluated from the total installation perspective considering contributions from the generator, the impedances of the power system and the transformer, and grounding connections. According to industry standards, total harmonic distortion (THD) on utility voltage should be limited to 5 percent and 3 percent for any single frequency. Current harmonics can be produced by rotating machines, inverters, or the saturation of transformers caused by 60-Hz overvoltages or DC arising from unbalanced inverters. Standards require DR suppliers to keep total harmonic distortion of current at the customer interface less than 5 percent. Table 4-3 shows the limits of harmonic distortion permitted for particular individual harmonics. Modern, force-commutated inverters tend to produce harmonics only greater than the 50th harmonic. However, some measurements show these may exceed levels of 1 percent.
Table 4-3 Harmonic Distortion Standards From IEEE 519-1992 Individual Harmonic 11
th th

Limit (%) 4

11 -17 17 -23
rd th

th

2 1.5 0.6 0.3 5

rd

23 -35
th

th

35 or greater Total Harmonic Distortion

The allowable aggregated rating of force-commutated inverters varies with the root-mean-square of the number of units applied because of phase cancellation among units. As an example, for a typical 4.16-kV feeder, 1, 9, and 100 inverters of similar sizes can have aggregated ratings of 0.5, 1.5, and 5.0 MW, respectively, and remain within harmonic voltage limits. At the 23-kV level allowable penetration exceeds 16 MWeven for a single unitwhich in practice makes the penetration unlimited.

4-64

The Impacts of Distributed Resources on Distribution Systems

Overvoltages
Overvoltages can be the result of load switching, variations in system generation, or variations in the reactive compensation of a system. Most equipment on a power system is only rated to withstand a voltage of 10% greater than rating for any length of time. Many types of sensitive loads can have even more stringent voltage requirements. Long-duration overvoltages must also be evaluated with respect to the long-time overvoltage capability of surge arresters. Metal oxide varistor (MOV) arresters in particular can experience overheating and failure because of high voltages of long duration. The most severe overvoltages that can occur can be caused by poor grounding configurations of DR sources and by ferro-resonance that may arise in some situations. Poor grounding and the overvoltages that occur during faults have been discussed in detail in the earlier section on transformers and grounding. Ferro-resonance comes in many forms but one common form is a resonance between system capacitances and non-linear inductances of transformer windings that arises in certain transformer and grounding configurations. Ungrounded, high-side transformer arrangements can be more susceptible to ferro-resonance than grounded configurations. Ferro-resonance can also occur with islanded induction generators that are self-excited by capacitors. Ferro-resonance can lead to distorted (harmonically rich) overvoltages that are up to several times the nominal level. Design approaches and operation practices exist that are known to help prevent ferro-resonance. DR sources need to be designed, grounded, and operated to minimize the danger of such overvoltages.

Flicker
Voltage flicker is a small but rapid fluctuation in system voltage of a few percent or less. The most commonly perceived result of voltage flicker is flickering output in lighting systems especially incandescent lights. Voltage flicker is caused by fluctuations in current flow on the power system. Such fluctuations may be the result of either DR or customer loads. In the case of DR, it may be caused by starting or stopping of generators, normal and abnormal output fluctuations of generators (misfiring reciprocating engines, photovoltaic or wind power variations), dynamic interactions with other loads or utility system equipment, and dynamic interactions with other DR units. If flicker becomes severe enough, it can be annoying to electric customers even to the point of causing severe distractions, including: loss of productivity, headaches, and other similar disturbances. Utility companies try to limit flicker so that it is at a level that cant be perceived by the human eye. This is accomplished by designing the power system to be sufficiently robust so that smaller load variations do not create noticeable voltage variations. It is also controlled by imposing limits on the types of loads that are allowed to connect at various points on the system. For example, many utilities have guidelines for the size of motors that can be connected at specific types of customers. To determine the appropriate levels of flicker that can be allowed on the system without causing a problem, many utilities have used (and continue to use) flicker sensitivity curves developed by 4-65

The Impacts of Distributed Resources on Distribution Systems

General Electric more than 50 years ago. These curves were developed by collecting data from test subjects (people) to determine their sensitivity to various levels of light flicker. Two curves have been used; these are the borderline or threshold of perception curve and the borderline or threshold of irritation curve (see Figure 4-37). Both of these curves are not flat but rather have a point of peak sensitivity that occurs in the 5 to 10-Hz region. In this region, the eye can detect a recurring voltage fluctuation of as small as approximately 0.25 percent. To limit voltage flicker to a level where it could not be detected by the eye, requires that the voltage fluctuations should not be greater than the indicated threshold of perception curve.

Figure 4-37 GE Flicker Curves

The IEEE has adopted the curves shown in Figure 4-37 as part of IEEE Standard 519-1992. However, more recently, the industry has moved toward a newer approach for flicker based on IEC standards7. The older GE approach was based on square fluctuations of voltage, however, the newer approach can handle just about any wave shape fluctuation and is considered more accurate and flexible. Even though the newer approach is more advanced and accurate, because
7

See IEC standards 61000-3-3, 61000-3-5, and 61000-3-7

4-66

The Impacts of Distributed Resources on Distribution Systems

of the long history of the GE flicker curves and its ease of application, it still is widely used today and considered quite adequate by many. The potential for a given DR to cause voltage flicker will depend on its size (capacity), type of technology, starting configuration, and the expected rate of fluctuating output current during typical operation. Each of the various types of DR has their own special characteristics that may contribute to flicker. For example, reciprocating-engine sets may be prone to flicker because the fluctuations that occur, because of misfiring cylinders, can be in the frequency range that is at the most sensitive point (5-10 Hz) of the flicker curve. Wind and photovoltaic systems dont fluctuate as rapidly but still can be rapid enough to create flicker. For example, cloud movements can cause localized photovoltaic arrays to produce sudden changes in real power equal to their rating and can do so in a matter of a few seconds in some cases. Wind turbines may also see fluctuations on the order of seconds up to minutes, as well because of wind variations. One type of wind turbine fluctuation involves the wind shading impact of the tower on the rotor and can result in more rapid, albeit smaller, fluctuation in output at a rate of 1-2 hertz. In addition, to these effects, the aggregation of units at different sites can increase the repetition frequency, but not the amplitude, of the associated voltage regulation in some cases. Another type of flicker that can occur is because of the starting of induction generators. If across the line starts are employed (the generator is brought up to speed with the power system connected), then a substantial voltage drop can occur on the system during the initial starting condition. The starting current could be anywhere from two to five times the full running current of the generator depending on the starting procedure employed. If this is done frequently enough then it becomes a flicker problem. For wind generators using induction machines, the units could connect many times per hour if the wind velocity was just hovering at the velocity threshold where the controller initiates the connection to the utility system. Speed matching before connection of the induction machine is one way to limit starting inrush. To screen applications like the reciprocating engines, PV, and wind systems for flicker, we need to identify the magnitude of the voltage fluctuations and the rate at which they occur. Then we can locate these coordinates on the flicker curve to see if they are visible. An easy way to quickly estimate the magnitude of fluctuations is to consider that for each 1 percent of the available, utility system, fault current that a generator injects into the interconnection point, the impact on system voltage is approximately 1 percent8. As an example, this means that a generator capable of generating a current equal to 5 percent of the utility system fault current at the point of connection would change the voltage by approximately 5 percent when it ramps up to full output. If the output of the generator in this example is fluctuating at a rate of 5 times per minute from off to full-on conditions, then it would
8

Note that this is the worst-case scenario based on injecting the current at the worst-case phase angle. In practical applications with the current injected at near unity power factor, the actual voltage change will be 0.5% to 0.9% for each percent of fault current injected.

4-67

The Impacts of Distributed Resources on Distribution Systems

create voltage flicker significantly greater than the threshold of irritation curve shown in Figure 4-37. This is a rather extreme example and most flicker is more subtle than this. A good screening tool to access the possibility of flicker is to divide the available utility system fault current at the point of interconnection by the DR rated output current. If this number is 400 or more, then the worst-case voltage variation could be no more than 0.25 percent, which, no matter how often the voltage fluctuates, should not be visible according to the GE flicker curve. This is an extremely conservative assessment and generally when the ratio of available fault current to DR rated output current is 100 or more, no visible flicker should exist. Any case, where the ratio is considerably less than 100 becomes a good candidate for potential flicker problems. Also, note that this screening did not consider starting of induction generators but rather steady state running variations of DR. Because of the high inrush associated with induction generator starting, a separate calculation would be needed to screen for flicker caused by that condition and it would consider the ratio of the utility system fault current to the induction generator starting current. Complex interactions occasionally can exist between DR and equipment on the utility system that can lead to flicker. For example, sub-synchronous power oscillations can exist between generators on a distribution feeder. Cyclic interactions can also exist between generators and loads. Generators and voltage regulation equipment may interact in a manner that causes repeated cycling of devices. These potentially can all cause flicker if left unchecked. For example, a generator with fluctuating output might cause a power-factor-correction capacitor to cycle on and off. These types of oscillations and interactions can be hard to predict and may arise unexpectedly as part of an installation. Once it is determined that voltage flicker may be an issue, many possible solutions can be employed to solve the problem. The most basic solution is to relocate the DR to a section of the feeder that has greater available fault current (it is stiffer). This approach could be feasible for utility-owned DR or merchant DR sites where the location of the site might be chosen from several possibilities. However, it is not highly feasible for customer-owned DR where there are no alternative sites. Other ways to reduce flicker include the use of DR controls that limit the rate of change of generator output or restrictions on the times at which generators are allowed to start and stop. For example, many inverters and rotational power generators include soft-start functions that allow them to slowly ramp up power over many seconds rather than step the power on. For starting of generators, precise speed matching for induction machines and tight synchronization at the moment of connection for synchronous machines will help reduce voltage perturbations caused during the utility system paralleling process. Use of a dedicated transformer will alleviate problems when flicker is confined to a secondary system with multiple customers. More advanced means of mitigating flicker may be required when a larger DR unit is applied on a feeder. Options include rapid-response voltage regulators (static VAR compensators) or fast-response reactive compensation using inverter reactive-power capabilities. Energy storage technologies can be applied to smooth the output fluctuations of solar and wind energy systems. Some of the mentioned solutions are quite costly and would normally be avoided where possible.

4-68

5
SUMMARY AND INTERCONNECTION SOLUTIONS
Previous chapters of this report have focused on the power system impacts and integration issues of DR. This final chapter focuses on the specific interconnection practices and practical approaches that can be used to determine requirements for DR installations. Some of the key elements associated with defining the specific interconnection requirements for DR that will be discussed in this section include: Screening of DR Standardized interconnection requirements (IEEE, state requirements, UL) Examples of typical interconnection layouts Relays used for DR Suitable voltage and frequency operating window Synchronization Protection for non-exporting DR Fast-acting reverse power protection DR communications

Screening of Potential DR Interconnections


Screening provides the opportunity to separate proposed DR interconnection projects into two bins: applications that need special attention and those that can go forward using a standardized template of interface requirements. The need for screening will become critical as the number of potential DR sites increases and as many smaller DR, such as residential fuel cells, microturbines, and PV systems, attempt to connect to the power system. Screening methodologies will allow utility engineers and planners to identify the interconnection requirements that are needed for safe, reliable, and efficient operation of each DR project with the need for fewer major system impact studies and less overall cost in the interconnectionevaluation process. Without screening, the existing process may not be efficient and streamlined enough to keep up with the influx of applications, and the industry may become quickly overloaded with DR interconnection applications. A typical screening process is shown in Figure 5-1. A key part of the process is collection of basic information describing the characteristics of the DR and the utility system at the planned point of interconnection. These data are used by the screening algorithm to identify specialneeds beyond the ordinary template of interconnection requirements. The algorithm evaluates the 5-1

Summary and Interconnection Solutions

system design to determine if technical problems will arise in a number of key areas, including voltage regulation, fault contributions, grounding, islanding, harmonics, power ratings, and other factors. The stiffness ratios that were discussed in Chapter 4 can play an important role in screening algorithms to help predict the size of the DR relative to power system capacity so as to predict its influence on the power system. Many different screening algorithms are being developed by industry and some states, such as Texas and California, have already formalized screening processes. The California screening process is shown in Figure 5-2.

Figure 5-1 The Basic Screening Approach behind all DR Screening Applications

5-2

Summary and Interconnection Solutions

Figure 5-2 Example of California Rule 21 Screening Process

Screening should consider the power system characteristics at the point of interconnection and not just the size of the DR. This is important because not all parts of the power system have the same impedance, so while one site can accept a large DR without an issue, another site may have serious system impact problems. To evaluate site sensitivity, a screening process can include a set of curves (or stiffness ratio calculations that accomplish the same effect) similar to those shown in Figure 5-3. This figure shows the likely maximum capacity of DR that a site can accept before system impacts force power system modifications to be made. To use the curves, simply identify the type of system to which the DR is to be connected and the proposed size of the DR and determine if it falls within the gray region. If the answer is yes, then it passes the screen. For example, the maximum size DR that could connect to a weak point on the system primary without a detailed study and potential modifications to the interface requirements would be about 50 kW according to these curves (note that these curves were developed for this report for illustration of this process and are not intended for actual screening). The preceding example was just one screen. To pass the full screening process, there would be many screening curves, calculations, and questions that would be addressed covering everything from the capacity of the system to the certification ratings of some of the key DR components. The screening tests should be packaged in a manner that is efficient (cost-effective to collect data and relatively fast to complete) but still is robust enough to meet the objectives of filtering out problematic installations.

5-3

Summary and Interconnection Solutions

Figure 5-3 Curves Identifying the System Impacts of DR Connecting to Various Points on the Distribution System

If the DR completely passes all screening tests, then it is assigned the standard interconnection requirements for its designated size class. If a DR application fails to pass the screening process, then it does not mean that the DR cannot be interconnected. It simply means that the DR is likely to have a significant impact and further study is needed to determine the scope of any problems and potential mitigation approaches. The types of studies that might be needed include load flow, short circuit, protection coordination, voltage flicker, harmonic, and stability studies. There would be three possible outcomes to the additional studies: 1. Analysis would show that the installation can proceed with the standard requirements (no changes at all to the basic interface design, essentially as though it had passed the screen in the first place). 2. Analysis would show that protection, control, and/or interface equipment changes made at the DR site could solve any issues, and these changes would be required to proceed. 3. Analysis would show that changes to the DR interface design alone are not enough. The utility system itself would need to be upgraded to handle the system impacts of the DR. The types of upgrades might involve the addition of regulators, new relays in the substation, additional protection devices, and so on.

Standardized Interconnection Requirements


In the preceding section on screening, standardized interconnection requirements were utilized if the DR could pass the screening approaches. However, the industry currently does not have 5-4

Summary and Interconnection Solutions

standardized requirements for all DR that are universally accepted. In the past two decades, there has been quite a bit of work to develop interconnection standards and guidelines for distributed generation. In the 1980s, utilities throughout the nation developed their own customized interconnection guidelines as a result of the PURPA legislation. These were the first generation of standards aimed at allowing independent power producers to connect to the system as a well as smaller-size DR installations. During this period, the IEEE and other organizations also developed a number of interconnection guidelines and standards such as IEEE 1001-1988, Guide For Interfacing Dispersed Storage and Generation to Facilities with Electric Utility Systems. This document is no longer in publication and is being superceded by new standards. Today, much progress has been made to facilitate the growing market for DR, and several standards are in place or in development. Just a few examples include: IEEE 929-2000, Recommended Practice for Utility Interface of PV Photovoltaic Systems UL 1741 Inverters, Converters and Controllers for Use in Independent Power Systems State Interconnection Requirements (NY, Texas, California, Vermont, other) Ongoing Updates of Existing Utility Interconnection Guidelines IEEE 1547, Standard for Interconnection of Distributed Resources with Electric Power Systems (currently in draft development stage)

The most significant standard underway related to distributed-generation is IEEE 1547. As of the writing of this report, it was at the Draft-8 stage and still undergoing further revisions prior to being completed. This standard is meant to provide a unified national standard for the interconnection of all distributed resources to the power system, and it is a necessary endeavor to complete it in order to better facilitate the screening process and streamline interconnection approaches discussed here. However, it is not clear in what form the final document will emerge because there is still much technical disagreement about the requirements and because the standard covers a very broad range of DR sizes (up to 10 MW). It has been suggested by many that the standard should be broken into more limited categories of size, such as DR less than 25 kW, DR 25 to 250 kW, and DR greater than 250 kW. This possible approach makes sense considering that the interconnection of a 10-MW generator is significantly different than the interconnection of a 5-kW residential PV system. Regardless, at this time there is not a unified national standard to fall back on for the screening process. Lacking this, the following sections discuss the general practices and approaches that are used and give some examples.

Interconnection Equipment at a Typical Site


Due to the lack of standard interconnections guidelines for most types of DG, most installations are based either on state guidelines or on existing utility company requirements around the country. Based on the types of installations today, Figure 5-4 describes the key elements of a typical generator interconnection to the primary feeder. These include the generator (with proper grounding), CT and PT to measure voltage and current conditions, the relaying protection package, low-voltage circuit breaker, disconnect switch, step-up transformer, and high-side protection device (fuse, recloser, or circuit breaker). There may also be a disconnect switch on the high side. Note the communication link, which is usually present only on larger systems. 5-5

Summary and Interconnection Solutions

Figure 5-4 Key Elements of an Interconnection

Most DR interfaces contain the generic components shown in Figure 5-4. However, there are substantial differences in the way the schemes are implemented based on the type of application, type of DR, electrical service layout, and size of the system. For example, small residential DR installations are typically single-phase, much less than 20 kW in rating, have no communication link, and operate at 120/240 volts. They also utilize the existing service transformer and fuse in most cases and be something like that shown in Figure 5-5. Note that in Figure 5-5, the power converter is an inverter with internal microprocessor relaying functions that perform the protection duties. Many of the residential DR of the future will be inverter-based due to the proliferation of fuel cells and PV systems.

5-6

Summary and Interconnection Solutions

Figure 5-5 Example of a Fuel Cell Interface at a Residential Location

Larger DR systems have more complex interconnections. The system of Figure 5-6 is a larger system composed of a synchronous rotating machine rated at several hundred kilowatts. It includes various discrete relays, more complex switchgear than the residential system, and communications/data logging. Figure 5-7 is a larger inverter-based system.

5-7

Summary and Interconnection Solutions

Figure 5-6 Interconnection of a 250-kVA Synchronous Generator (Alternate Transformer Shown Is for Three-Wire Ungrounded Feeders)

5-8

Summary and Interconnection Solutions

Figure 5-7 Example of an Interconnection for a Large Three-Phase Inverter

A comparison of the different scales of DR installation and the characteristics associated with the interconnection hardware are shown in Table 5-1. Note that many of the basic elements are the same from a functional perspective regardless of size but that they are implemented much differently depending on the scale. For example, the grounding of the generator becomes a much more complex issue for larger three-phase systems than small single-phase systems, as do the protection relay packages and switchgear arrangements.

5-9

Summary and Interconnection Solutions


Table 5-1 Comparison of Typical Interconnection Equipment Used in DR Applications
Small Scale Residential DR Usually single phase, less than 20 kW, 120/240 V-typically a single unit Use existing primary fuse Small Scale Commercial and Industrial DR Usually 25-100 kW, three phase, 208Y/120 or 480Y/277 V typically a single unit Intermediate Scale Commercial and Industrial DR Usually 100-1000 kW, three phase, 208Y/120 or 480Y/277, volts often multiple units in parallel 3 pole switch and/or breaker may need to be added Larger Commercial and Industrial DR >1000 kW, threephase, 480Y/277 volts (may be higher) usually multiple units in parallel 3 pole switch and/or breaker may need to be added Would likely use existing site transformer if grounding is appropriate there can be a second cascaded isolation transformer must be very careful about utility system grounding needs Yes - if generation is LV but generators may be HV, so HV switchgear may be employed instead

Description

Type of DR size, and operating voltage Primary protection device and disconnect means

Depends on layout of facility

Interface and step-up transformer

Existing distribution transformer for customer is almost always suitable

Would likely use existing site transformer consideration of utility grounding needs required

Would likely use existing site transformer if grounding is appropriate there can be a second cascaded isolation transformer must be very careful about utility system grounding needs

Low voltage disconnect switch for generator

Yes (must be visible break, load break, lockable disconnect switch) Fuse or breaker suitable Basic protection as per IEEE 929/UL1741 for inverters and basic anti-island protection for rotating equipment

Yes (must be visible break, load break, lockable disconnect switch) Breaker typical molded case circuit breaker unit may also use contactor in addition

Yes (must be visible break, load break, lockable disconnect switch)

Low voltage protection device

Likely motor-operated circuit breaker

Breaker - if generators are LV

Generator and utility system protection relays

Fairly intensive protection but may be lacking some functions of the larger generators

Fairly intensive protection but may be lacking some functions of the larger generators

Full protection package similar to a large central station generator

Grounding

Simple single phase grounding arrangement is simple ground per NEC code

3-phase grounding requires assessment of generator fault capabilities careful review of transformer type and utility system requirements No (extremely rare if ever)

3-phase grounding requires assessment of generator fault capabilities careful review of transformer type and utility system requirements Transfer trip occasionally employed

3-phase grounding requires assessment of generator fault capabilities careful review of transformer type and utility system requirements Transfer trip often used for units this large

Remote trip/control functions

No (never)

5-10

Summary and Interconnection Solutions

Relays for Protection of Synchronous and Induction Generators


There has been much focus in recent years on the design specifications for the types of protective relays used for interfacing rotating distributed generators to electric power systems. Not all relays are equal in quality and performance. A crude undervoltage relay suitable to protect a small motor would not be suitable for protecting a distributed generator due to limitations in its accuracy and other performance factors. Relays vary in quality and sophistication from the most basic commercial-grade relays up to high-performance utility-grade relays. The power system infrastructure including central station generators, transmission lines, distribution lines, and substations employ utility-grade relays for protection. As a result, electric utilities have gained much comfort and experience in working with such relays due to their very high accuracy, security, reliability, and ease of testing. While there is no set definition of a utility-grade relay, generally it is a relay that has the following characteristics: Meets or exceeds ANSI/IEEE standards for surge withstand and other operating characteristics High accuracy in pickup and time settings Positive indication of the trip cause (indicated on the display) Draw-out mounting Test signal inputs for in-place testing High quality design and high reliability/security Detailed documentation covering application, testing, maintenance, and service Can be either an electromechanical-based (older design) or microprocessor-based design (newer design)

All early utility relays were electromechanical relays. Much protection today on the power system is still performed by such relays, and they are still commercially available. However, over the past 10 to 15 years, microprocessor-controlled relays have come of age and have become widely used. It is expected that they will eventually replace most of the electromechanical relays used in utility-scale applications. Microprocessor relays are more adaptable and flexible because they can be programmed to emulate a wide range of electromechanical protection functions and can act in ways not possible with electromechanical systems. Some of the newer microprocessor relays include waveform-recording features so that the event that tripped the unit can be digitally recorded. Both the microprocessor relays and the older electromechanical designs are available in utility-grade packages (see Figure 5-8).

5-11

Summary and Interconnection Solutions

Figure 5-8 Utility-Grade Relays (Photo Courtesy of GE)

With the older electromechanical relays, an entire rack of relays would be needed to protect a generator because generally these types only offer one function per relay and sometimes 10 or more relays would be needed to protect a generator (considering all of the functions that are needed and that there are three phases and a neutral to protect). The cost of a full electromechanical relay set to protect a single generator could easily exceed $20k when all of the ancillary components are factored in. Modern microprocessor-based relays can be multifunctional, so a single multifunction relay can provide most or all of the functions needed to protect a generator, including under- and over-voltage, under- and over-frequency, overcurrent, and unbalance. Several manufacturers offer utility-grade multifunction relays aimed specifically at the protection of small generators for a few thousand dollars, and this certainly represents a considerable improvement in cost/functionality compared with the older single-function relays. As a cost saving measure, many DR installers are promoting the use of commercial/industrialgrade relays that are of lesser quality than the utility grade counterparts. But electric utilities have resisted this usage because there is valid concern that the industrial relays are not sufficiently accurate and reliable to meet the requirements of protecting DR and do not have suitable test ports needed to verify settings. These industrial-grade relays were really just intended as motor protectors for under-voltage, unbalance, and so on, and are not really intended for generator protection in a demanding utility connected application. As newer multifunction microprocessor relays have become available that are utility-grade and far less costly than their predecessors, the argument for using anything other than a utility-grade package on rotating generators does not really make sense, especially for generators greater than 500 kW where the cost of the multifunction relay is less than 2% of the system cost. For smaller rotating machines (much less than 500 kW), relay cost does start to become a significant portion 5-12

Summary and Interconnection Solutions

of the system cost. However, most smaller DG will be inverter-based, which do not need a separate add-on relay package (see next section).

Relays for Protection of Inverters


Inverters are a very different situation than rotating machines. Inverters, by the nature of the way they convert DC power to 60 Hz, have precise control of their power production and very highspeed response to abnormal system conditions. As a result, a well-designed microprocessor control that is part of the inverter itself can perform better at protecting the inverter and utility system than a discrete add-on relay. It has generally become accepted within most technical circles that small inverters can successfully utilize their own internal protection functions in lieu of add-on relays as long as the inverter is type tested as specified in various industry guidelines. These include tests such as those discussed in UL1741 and IEEE 929-2000 for nonislanding inverters. Some states such as New York and Texas have adopted language in their interconnection guidelines that specify that type-tested inverters can be directly connected to the system if they pass the test. An example of an inverter with built-in microprocessor control functions, that should ordinarily need no additional relays, is shown in Figure 5-9.

Figure 5-9 Example of an Inverter Product Intended for Photovoltaic Applications That Has Internal Microprocessor-Based Protective Functions (Photo Courtesy of Xantrex)

5-13

Summary and Interconnection Solutions

While separate add-on relays may not be required for most small inverters, there are situations where such relays might be employed. In the case of very large inverter systems (several hundred kilowatts or larger), such relays might be used as backup over the inverter controls. There is also the practical situation where the inverter feeds into a facility bus, and the utility system protection point (where isolation will occur) is at a different location than the inverter (see Figure 5-10). For that situation, a separate package of relays might be employed to control the main connection to the utility system. This arrangement also allows the generation feeding the bus to be configured to carry the site load as an island while the entire site remains disconnected from the utility system.

Figure 5-10 When the Inverter Feeds a Bus at a Different Location Than the Intended Interconnection Isolation Point, a Separate Control Package May Be Needed for That Point

Relay Functions
The American National Standards Institute (ANSI) and IEEE have adopted designating device numbers to indicate the protective functions on system design drawings. On such drawings, protective relays are described as a circle with a number inside that designates the associated function. The key functions that are used in system protection are described in Table 5-2. Note that these functions may be implemented in as many single function relays, a multifunction relay, or internally within a microprocessor-controlled inverter. The implementation simply depends on the type of DR, its application, and the applicable rules governing the protection requirements. 5-14

Summary and Interconnection Solutions


Table 5-2 Protection Functions Commonly Employed On Larger Generators

Protection Objective
Detect Islanding

Function Used
81o/u (under or over frequency) 81R (rate of change of frequency 27/59 (Under and over voltage) Transfer trips Phase faults 51V, 67, 21 Ground Faults 51N, 67N, 59N, 27N 46 (unbalanced current) 47 (unbalanced voltage)

Comments
Active anti-islanding may also be employed on inverters 51V is voltage restrained overcurrent that is used to increase sensitivity These may cause overheating of DG If used looking toward generator, generator prevents motoring. If used looking towards utility system, it prevents export of power Power.

Detect Faults and Overloads Detect Potentially Damaging Unbalances to Generator Detect Abnormal Power Flow

32 (reverse power)

Allow Machine machine to resynchronize Resynchronize

25 (synchronizing relay)

Suitable Voltage and Frequency Operating Window


Generators connected in parallel with the power system should be operated only when the voltage is within a suitable operating window. This is to ensure that the generator does not operate in a sustained manner at levels of voltage or frequency that may be damaging or disruptive to loads and equipment. The suitable operating window also ensures that abnormal voltages due to faults can be detected and used for anti-island protection purposes. Voltage and frequency relays on rotating generators and internal microprocessor controllers on inverters are used to monitor under- or over-voltage and under- or over-frequency. These protective functions disconnect (or de-energize) the DR upon detection of the out-of-range condition. The IEEE and UL have defined the suitable operating limits for distributed generation in published standards such as IEEE 929-2000 and UL1741. The IEEE 1547 draft standard discusses acceptable voltage operating limits for distributed resources (see Table 5-3). For frequency, the window that the industry is focusing on as being most appropriate is a fixed window of 59.3 Hz to 60.5 Hz for DR less than 10 kW and an adjustable window for larger DR to coordinate with under-frequency load-shedding schemes.

5-15

Summary and Interconnection Solutions


Table 5-3 Recommended Voltage Operating Limits for Small Distributed Resources (< 30 kW) as Described in IEEE 1547 Draft 8

Voltage Range (% of base Note 1 voltage) V<50 50V<88 88V<110 110V<120 V= 120

Clearing Time (seconds) 0.16 2 Operation Allowed (no trip) 1 0.16


Note 2

Note 1: Base voltages are the nominal voltages stated in ANSI C84.1 Note 2: DR30 kW Maximum Clearing Time. DR>30 kW, Default Clearing Time or area EPS operator may specify different voltage settings or trip times to accommodate area EPS system requirements

Note that settings shown in Table 5-3 are still being debated, so minor changes from these settings may occur in the final published IEEE 1547 document. Nonetheless, these are representative of the types of settings that would be typically employed on generators.

Synchronization
Synchronous generators and self-commutating inverter sources should be synchronized with the utility system prior to parallel operation with the power system. Perfect synchronization of a generator upon interconnection is defined as having the exact frequency, phase angle, and voltage magnitude as the utility system voltage at the moment of connection. With perfect synchronization, there is no voltage disturbance imposed on the system at the moment of connection. Perfect synchronization cannot be achieved due to limitations in the ability to control the generator and accuracy limits of the relays. In practical applications, there is an acceptable window of synchronization where the voltage and frequency are close enough to the utility system that the difference will not cause any significant problems at the moment of connection. The IEEE P1547 Working Group is developing synchronization requirements for generation. The limits are still in a draft form and could change but were as shown in Table 5-4 at the time that this report was prepared. These recommended settings are reasonable from the perspective of their achievability (cost and accuracy of available control equipment) and suitability for limiting negative impacts on the power system at the moment of interconnection.

5-16

Summary and Interconnection Solutions


Table 5-4 Synchronization Requirements of DR at Moment of Connection

Aggregate Rating of DR Units


(kVA) 0-500 >500-1,500 >1,500-10,000

Frequency Difference
(Df, Hz) 0.3 0.2 0.1

Voltage Difference
(DV, %) 10 5 3

Phase Angle Difference


(Df, degrees) 20 15 10

Note that for synchronous rotating generators, the generator physically remains disconnected from the line until the conditions of Table 5-4 are satisfied. The generator connects to the system by the closing of a mechanical contactor, switch, or circuit breaker. In the case of inverters, a similar process can occur, but depending on the size and design of the inverter, the connection may be performed either electronically via static devices or mechanically.

Non-Exporting DR as an Alternative to Employing Protection Functions


It is sometimes suggested that DR that do not export power into the utility system under steadystate conditions will not impact the bulk power system in any significant way. As a result, nonexporting generators may operate safely with less protection than exporting generators. While there is validity that certain power system impacts are less for non-exporters (such as effect on voltage, losses, and harmonics), many of the potential impacts, especially related to fault-current contributions and islanding on the utility system, can still occur even if generation is operated as non-exporting in the steady state. For example, a DR adjusted to operate as non-exporting under steady-state conditions when exposed to the sudden serious voltage drop on the power system caused by a fault can contribute fault current until the unit is tripped. Non-exporters can also still island with the power system under the right circumstances and must have a means of anti-island protection. They also must synchronize before connecting and will therefore need synchronization protection. Overall, non-exporters still must be fully protected for many scenarios. One big difference for a non-exporting DR is that a reverse power and/or directional overcurrent relay can be used as an anti-island tool quite effectively in lieu of voltage and frequency relays. However, to ensure that a non-exporter is not operating outside the normal voltage and frequency limits, it still would likely employ voltage and frequency protection anyway. So, overall, non-exporters still end up with pretty much the same array of relays needed for exporting systems.

5-17

Summary and Interconnection Solutions

Fast-Acting Directional Protection


A fast acting directional overcurrent device that can quickly trip the generator and stop the fault contributions into the distribution system can be effectively deployed to help reduce the impact of DR on the system. This type of protection would generally be deployed only for one-way (non-export) generators. There are several options available (see Figure 5-11). The first option is a reverse power relay (known as device 32) that can be set to trip on a small amount of reverse flow back into the utility system. A second option is an instantaneous directional overcurrent relay (known as a function 67). Both of these would be used to control a mechanical interrupting device such as a circuit breaker. For these relays, about 1 to 2 cycles are required to operate, and then the mechanical breaker requires another 1.5 to 3.5 cycles to interrupt the fault current. The total response time of the system is 2.5 to 5 cycles. This option will limit the duration of the fault contributions to 2.5 to 5 cycles and therefore has a positive impact on limiting the influence of the DR fault contributions on the slower-acting utility system protection devices, such as timedelayed circuit breakers and reclosers. A third device that is even faster than any of the relay/breaker options is a static switch. Static switches can interrupt current in cycle if reverse flow is detected. This would consist of an electronic monitoring circuit that triggers a silicon-controlled rectifier (SCR) or transistor switch. This switch is just like those used in UPS systems and could limit the fault contributions of the DR to about cycles in duration. With this level of speed, the influence of the DR fault-current contributions on utility system equipment is essentially eliminated, and even the fastest-acting circuit breakers on the utility system will not be influenced by the fault contributions. Use of a static switch also allows for arrangements that can provide high power quality to be maintained at the DG site during feeder disturbances.

Figure 5-11 Two Methods of Relatively Fast-Acting Reverse Flow Protection

While the use of fast-acting protection schemes to detect and prevent flow of reverse current into the power system has been employed for many DR applications, such schemes are not without significant drawbacks. From the perspective of DR, the most significant is that many fast-acting schemes would experience nuisance trips on an almost daily basis, given the power quality conditions that occur on distribution systems. Nuisance trips cause increased fatigue on system components because the system would constantly be dealing with large load rejections. Nuisance 5-18

Summary and Interconnection Solutions

trips might also interfere with sensitive loads and combined heat and power applications where the generator needs to run constantly to maintain stable heat conditions. The usual solution to reduce nuisance trips is to desensitize the protection settings by adding a time delay or increase the trip threshold. Common schemes would employ a time delay of at least 10 cycles (and sometimes much longer). However, when time-delay protection is employed, the whole benefit of using the fast-acting reverse power trip disappears from the perspective of the fault contributions to the utility system. If a DR protection device could be developed that never exported fault contributions longer than cycles in duration on a reliable basis without nuisance trips, then it would be a significant breakthrough in DR protection. Table 5-5 summarizes the value of different reverse flow protection schemes at mitigating fault contributions.
Table 5-5 Comparison of the Benefits of Various Reverse Current Tripping Techniques from the Perspective of the Impacts of Generator Fault Contributions on Utility System Protection Devices Speed of Response of Tripping the DR when Utility System Fault Occurs Advantages/Drawbacks

Fast Acting Static Switch Device - Fault Contribution Limited to cycle or less

Solves nearly all coordination problems with utility system circuit breakers, reclosers, fuses and other devices. But it is prone to nuisance trips, costs more than a mechanical breaker, is less reliable and wastes about 1% of the energy passing through it. Solves coordination issues only with time-delayed devices. Instantaneous trip devices and some fuses are still impacted. It is prone to nuisance trips much more than a time-delayed reverse power relay but less than the static switch. Costs much less that a static device, is more reliable and more efficient than static device. The time delay may solve most nuisance trips, but the scheme is not fast enough to offer any benefits from the perspective of mitigating fault contributions.

Fast Acting Directional Overcurrent Relay (67) or Reverse Power Relay (32) Controlling a Circuit Breaker - Fault Contribution Limited to 2 to 5 cycles

Reverse Power or Directional Overcurrent Relay with 30 Cycle Time Delay Controlling a Circuit Breaker Fault Contribution Limited to 30 cycles

Communication
Communication for DR is most commonly employed for a transfer-trip function. The most common form of transfer trip is to send a signal from a distribution system protective device, such as a feeder circuit breaker in a substation, out to a distributed generator (see Figure 5-12). With this arrangement, when the circuit breaker in the substation trips, a signal is sent instantly to the DR unit to trip at the same time.

5-19

Summary and Interconnection Solutions

Figure 5-12 Example of Transfer Trip Used to Ensure that the DR Is No Longer Operating

The value of transfer trip is that it ensures that the DR will be tripped offline whenever the feeder circuit breaker opens, thus avoiding the danger that the DR will keep the feeder energized causing an unintentional island. It may seem redundant to have a transfer trip when all generators are required to have autonomous anti-island protection by means of voltage and frequency relays. However, it is important to recognize that the anti-island protection is not 100% reliable, so the transfer trip ensures that the generator will be tripped. Many utilities require a transfer trip for the larger DR (> 1000 kW), and some even require it for relatively small DR down to a few hundred kilowatts. Transfer trip is very expensive because it needs a fast-responding dedicated communication link. This is usually done with dedicated telephone lease line circuits. The annualized cost of a transfer trip, including equipment and communication costs, can easily exceed $10,000 per year, so it is prohibitively expensive for smaller distributed generators. Other communication technologies are being considered as lower-cost alternatives, including the Internet and various wireless and power-line carrier technologies. These may cost much less than $100 per year, especially when implemented in piggyback with other services, but they will need to be set up properly to ensure sufficient speed and reliability comparable to those of a lease line. In the future, these communications will be very important to the commercial success of DR because low-cost and reliable communication will be a key element of the system control and coordination. It is not likely that high penetration scenarios of DG predicted in the future can be achieved successfully without a more widespread adoption of communication for even the smaller DG. But for this to happen, costs need to come down from what they are for current transfer-trip technologies.

Overall Conclusions on the Complexity of Interconnection


It has been the intent of this report to serve as a tool to assist utility planners, energy providers, regulators, and other DR stakeholders in formulating the appropriate policies and technical requirements that balance the needs of distributed generation with the safety and reliability needs 5-20

Summary and Interconnection Solutions

of electric power systems. A wide range of issues related to DR integration have been covered, and from these discussions it is clear that there are safety and power system impact issues that must be considered when interconnecting DR to the power system. Nonetheless, small to modest penetrations of DR can be successfully integrated with the power system as long as the interconnection designs meet the basic requirements that consider the issues of safety, reliability, power quality, and system efficiency. The existing distribution system design, while not ideal for DR, has the capability to handle the expected modest amounts of DR during the near term. The new screening procedures that are under development and the interconnection standards that are soon to be available should also support this process. Looking farther into the future, however, if a high level of DR penetration occurs, as is predicted by many, then much work remains to be done to develop the control equipment, more compatible power system designs, operating procedures, and communication methodologies that are needed to operate a DR-rich power system. While the DR-intense operating environment could still be 10 or 20 years away, now is the time to begin the process of developing strategies that will allow this path, if it is viable, to be available as an option for the future.

5-21

Target: Distributed Resources: Information for Business Strategies

About EPRI EPRI creates science and technology solutions for the global energy and energy services industry. U.S. electric utilities established the Electric Power Research Institute in 1973 as a nonprofit research consortium for the benefit of utility members, their customers, and society. Now known simply as EPRI, the company provides a wide range of innovative products and services to more than 1000 energyrelated organizations in 40 countries. EPRIs multidisciplinary team of scientists and engineers draws on a worldwide network of technical and business expertise to help solve todays toughest energy and environmental problems. EPRI. Electrify the World

2001 Electric Power Research Institute (EPRI), Inc. All rights reserved. Electric Power Research Institute and EPRI are registered service marks of the Electric Power Research Institute, Inc. EPRI. ELECTRIFY THE WORLD is a service mark of the Electric Power Research Institute, Inc. Printed on recycled paper in the United States of America 1004061

EPRI 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10412, Palo Alto, California 94303 USA 800.313.3774 650.855.2121 askepri@epri.com www.epri.com

Das könnte Ihnen auch gefallen